Graphical Abstract Figure
Graphical Abstract Figure
Close modal

Abstract

Regarding high-precision temperature field prediction of a heating ventilation and air conditioning (HVAC) unit using thermal fluid dynamics simulation, the determination of the turbulent Prandtl number (Prt) was a key issue. In this study, we present an attempt to improve the accuracy of thermal fluid dynamics simulations in HVAC units by adjusting the Prt using data assimilation techniques. First, we simultaneously measured the velocity and temperature in the hot and cold air mixing zone of a simple HVAC model using particle image velocimetry and thermocouples. Second, we coupled data assimilation to the thermal fluid simulation model to determine the Prt in the mixing field. Finally, we proposed two functions of the Prt for high velocity and low velocity regions using multidimensional analysis. In the future, we believe that the Prt functions can be applied to thermal fluid simulations of actual HVAC unit to accurately predict performance without conducting prototype experiments, thereby contributing to reducing development cost and time of HVAC unit.

1 Introduction

A heating ventilation and air conditioning (HVAC) unit is an essential unit to adjust temperature for passenger's comfortability in an automotive cabin as shown in Fig. 1, dehumidifying/cooling and heating the air with the evaporator and the heater core inside the HVAC unit, and then adjusting the temperature with the airflow control door and switching the air outlet. In recent years, due to high performance requirements of automotive air conditioning systems, the complexity of HVAC internal flow path is progressing, i.e., left/right independent process control type for automobile, rear air conditioning for the rear seat passengers, etc., understanding the heat transfer phenomenon inside the HVAC unit becomes more and more important in HVAC unit design [13]. To achieve the demand target in performance improvement at an early stage of development, it is indispensable to grasp the information on flow and temperature field using thermal fluid dynamics (CFD) simulation. However, if a default setting of the commercial CFD simulation software is utilized to predict temperature at duct outlet of HVAC unit, in which Prandtl number is usually set to about 0.7–0.9 as laminar flow, there is a big difference between simulation results and the measurement data, because the airflow inside HVAC unit has a very complex turbulence vortex structure and the structure of the turbulence airflow changes with the variation of the gap between the airflow control door and the associated shut rib, due to the various opening degrees of the control door. Therefore, as for high-accuracy prediction of HVAC performance at an initial stage of the development using CFD simulation, the determination of the turbulent Prandtl number (Prt) in the CFD model is a key issue [4,5].

Fig. 1
The HVAC unit is used in an automotive air conditioning system
Fig. 1
The HVAC unit is used in an automotive air conditioning system
Close modal

The Prt, which represents the dissimilarity between turbulent momentum and heat transfer, is useful for solving the heat transfer problem of turbulent flow. So far, based on Reynolds' analogy, many experimental and theoretical investigations have been carried out for determination of the Prt; Blom [6] experimentally determined the Prt in a development temperature boundary layer, Reynolds [7] examined more than 30 ways of predicting the Prt, Browne and Antonia [8] measured the Prt in the self-preserving region of a slightly heated plane jet, Kays [9] examined the experimental data on the Prt for the two-dimensional turbulent boundary layer and for fully developed flow in a circular tube or a flat duct, Morris et al. [10] investigated the role of four alternative turbulent Prandtl number functions on the predicted heat transfer coefficients obtained by numerical analysis, Churchill [11] presented a detailed interpretation of the Prt, Redjem-Saad et al. [12] investigated the Prt by a direct numerical simulation (DNS) of heat transfer in a fully developed turbulent pipe flow, Venayagamoorthy and Stretch [13] derived a general relationship for the Prt for homogeneous stably stratified turbulence from the turbulent kinetic energy and scalar variance equations, Li [14] theoretically investigated the Prt in the atmospheric boundary layer, Basu and Holtslag [15] proposed a simple analytical approach based on the variance and flux budget equations to quantitatively evaluate the stability dependence of the Prt, Xu et al. [16] investigated the behavior of the Prt for buoyancy-affected flows near a vertical surface using machine learning techniques based on DNS data base, etc. However, due to theoretical deficiencies and lack of measurement precision, the results showed that qualitative and quantitative differences were too large to draw conclusive conclusions about the general situation. Fortunately, several pieces of evidence have shown that the Prt which should be a function of local turbulence properties is often assumed in many engineering calculations to be a function of velocity and temperature, although a clear variation characteristic of the Prt in the practical range has not yet been clarified.

In this study, we endeavor to improve the accuracy of the temperature field prediction in turbulent mixing for the HVAC unit, tuning up model parameter (the Prt) in the CFD simulation by functionalization and data assimilation with actual measurement data, and the effects are compared and discussed. First, to understand the mixing process of hot and cold air in HVAC unit, we carry out an experimental study on mixing field of a simple HVAC model, and the velocity and temperature fields of mixing zone are simultaneously measured using particle image velocimetry (PIV) and thermocouples. Second, to compensate for the deficiency that the spatial and temporal measurement resolution of the temperature field is lacking compared to velocity field measurements, data assimilation is performed based on these experiments that simulate the mixing field inside the simple HVAC model and the Prt distribution that fits the experimental data which are predicted. Finally, to employ the predicted Prt in the thermal fluid simulation of the actual HVAC unit, we use multidimensional analysis to derive two functions of the Prt for high velocity and low velocity regions.

2 Experimental Method

We experimentally create several hot and cold air mixing fields with different velocity and temperature conditions in a simple HVAC model, and measure the air velocity and temperature of these mixed fields using particle image velocimetry and thermocouples, respectively.

Figure 2 shows a schematic diagram of experimental apparatus: two dryers with multi-stage speed and temperature adjustment functions are installed between the smoke chamber and the simple HVAC model. The outside air drawn into the smoke chamber is split into two air streams, which are heated and cooled by each dryer. After passing through each rectifier, the hot and cold air streams are mixed inside the simple HVAC model. By adjusting the velocity and temperature of the blowing air with the two dryers, it is possible to reproduce the various mixing field that are expected in an actual HVAC unit.

Fig. 2
Schematic diagram of experimental apparatus
Fig. 2
Schematic diagram of experimental apparatus
Close modal

The simple HVAC model is made of transparent acrylic plate (thickness 5 mm) as shown in Fig. 3, and two guide plates are installed instead of the control door in the actual HVAC unit. In this experiment, the smoke of incense sticks is used to visualize the velocity field by PIV, and the temperature data of 60 points in the air mixing zone of the simple HVAC are measured by thermocouples, where the diameter of thermocouple is 0.05 mm, and be fixed on the pins with a height of 25 mm. However, if the temperature data of 60 observation points are measured at the same time, it is concerned that the pin in the upstream disturbs the downstream and the physical quantity in the mixing field cannot be observed correctly. Therefore, under a predetermined experimental condition, only one row of 10 pins is installed in the x direction, and multipoint temperature measurement is performed six times with moving the position of pins from the front to the rear in the y direction.

Fig. 3
A simple HVAC model
Fig. 3
A simple HVAC model
Close modal

Figure 4 displays an overview of the experimental apparatus: the PIV system consists of a high speed camera (Photron, Fastcam SA-X2) equipped with Nikkor 50 mm f/1.2 that has a narrow band bandpass filter to avoid outside light, a double cavity Nd:YLF laser (Photonics Industries, DM30-527DH), and an analysis software (Dantec Dynamics, Dynamic Studio 6.1). The pair of particle images is captured at 2 kHz by using the frame-straddling method and instantaneous velocity field is calculated with adaptive PIV algorithm. Mean velocity fields and statics are calculated from 3638 instantaneous velocity files.

Fig. 4
Overview of the experimental apparatus
Fig. 4
Overview of the experimental apparatus
Close modal

In this experiment, to create as many variations in the velocity and temperature as possible, considering the range of velocity and temperature inside the actual HVAC unit (0–15 m/s and 5–85 °C), the experimental conditions are set using multi-stage adjustment function of the dryer, and these specific values are obtained by anemometer and thermocouple measurements at inlets of the simple HVAC model as shown in Table 1. The surroundings temperature and relative humidity were maintained at 20 °C and 40%, respectively.

Table 1

Experimental conditions of inlet air flow

No.Low temperature airHigh temperature air
Tl (°C)Ul (m/s)Th (°C)Uh (m/s)
Test 0112.14.5187.43.47
Test 0212.46.0684.74.90
Test 0313.57.9781.56.11
Test 0411.77.7978.52.64
Test 0544.97.9860.55.96
Test 0631.77.8066.76.02
No.Low temperature airHigh temperature air
Tl (°C)Ul (m/s)Th (°C)Uh (m/s)
Test 0112.14.5187.43.47
Test 0212.46.0684.74.90
Test 0313.57.9781.56.11
Test 0411.77.7978.52.64
Test 0544.97.9860.55.96
Test 0631.77.8066.76.02

3 Experimental Results and Discussion

3.1 Error Estimation of Measurement.

As a simultaneous measurement of the velocity by PIV and the temperature by thermocouple, it is necessary to discuss three influences on the flow field, such as the particle size, the particle concentration, and the temperature detection part of thermocouple. First, since the particle size in incense smoke is small enough (approximately less 1 μm), the influence of the particle size can be omitted [17]. Next, the fans in the two dryers diffuse the incense smoke to produce uniform smoke in this experiment, and we also carry out a preliminary experiment to determine the number of incense sticks in the chamber so that the variation in smoke concentration have no effects on measurement, so the influence of the particle concentration can also be ignored. Last, in order to avoid the influence from the temperature detection part of thermocouple, with adjusting the installation position of laser light sheet, the temperature detection part of thermocouple is fixed as close as possible to the measurement surface of PIV without disturbing the flow field; Fig. 5 compares the measured data of the velocity component without thermocouples and that with thermocouples under experimental conditions of test 04; it shows that thermocouples have almost no effect on the flow field of PIV measurement.

Fig. 5
Comparison of the velocity component without thermocouples and that with thermocouples
Fig. 5
Comparison of the velocity component without thermocouples and that with thermocouples
Close modal
For the temperature measurement of fluid flows, the heat conduction along the temperature sensor body causes the measurement error [18], and to correct this measurement error, in consideration of the thermal balance between the thermocouple wire and the surrounding fluid, the heat conduction error in fluid temperature measurement is theoretically analyzed based on the following equation,
(1)
where Tf is the temperature of fluid, Tw is the temperature of sensor, ρw(=8930kg/m3) is the linear density of wire, rw(=0.025mm) is the radius of wire, cw(=410kJ/(kgK)) is the heat capacity of wire, τ(ρwcwTw/2h) is the thermal time constant, and the heat transfer coefficient h is estimated by Kramers equation:
(2)
where kf is the thermal conductivity of fluid, Reynold number Ref and Prandlt number Prf are calculated from the average measurement data of fluid.

Figure 6 shows the error of the measured value and the corrected value under the experimental conditions of test 02; the maximum relative error and the absolute error are less than 2.0% and below 0.8 °C, respectively. However, the corrected value of temperature is used in the research.

Fig. 6
The error of the measured value and the corrected value
Fig. 6
The error of the measured value and the corrected value
Close modal

3.2 Characteristics of Mean Field.

Hereafter, only the results in which remarkable differences are observed under each of the experimental conditions are described. We select the case of high velocity (test 03) and the case of high velocity difference (test 04) with large temperature difference between the hot air and the cold air.

Figure 7 shows the mean velocity field and the mean temperature field in case of test 03 and test 04. Although the temperature difference between the hot air and the cold air is almost the same, the temperature fields are mixed well in the case of test 03, the reason is that a stronger vortex distribution is generated in the high velocity field of test 03 and the streamwise vortex promotes mixing as shown in Fig. 8, where the dimensionless vorticity is defined as ωzd/Ul using the distance between thermocouples d and the velocity of cold air Ul. As a result, by damming the flow with guide plates, a strong exfoliation shear layer is formed, and it also appears that the mixing is facilitated by mechanisms such as Kelvin–Helmholtz instability.

Fig. 7
The velocity field and temperature field of test 03 and test 04
Fig. 7
The velocity field and temperature field of test 03 and test 04
Close modal
Fig. 8
The dimensionless magnitudes of the streamwise vorticity
Fig. 8
The dimensionless magnitudes of the streamwise vorticity
Close modal

3.3 Fluctuation Terms of Turbulence Field.

The velocity field of turbulent flow can be split into a mean term and a fluctuating term using Reynolds decomposition; the Reynold stresses component uv¯ reflects the fluctuating velocity intensity distribution in the turbulent field as shown in Fig. 9; it shows that the fluctuation strength in the case of test 03 is larger than that in the case of test 04, and especially in the area between the two guide plates.

Fig. 9
Reynolds stresses component u′v′¯
Fig. 9
Reynolds stresses component u′v′¯
Close modal

Since the heat transfer between the hot air and the cold air predominantly occurred in the direction perpendicular to the mainstream direction, a local coordinate (xi*,yi*) is introduced as shown in Fig. 10 

Fig. 10
A local coordinate (xi*,yi*)
Fig. 10
A local coordinate (xi*,yi*)
Close modal
(3a)
(3b)

Figures 11(a) and 11(b) plot the turbulent heat flux v*T¯ and the average temperature gradient in the direction perpendicular to the mainstream direction in case of test 03 and test 04, respectively. In comparison with two distributions, it is hard to find a clean relationship with turbulent heat transfer and temperature gradient, but both of these maximum intensity values appeared in the mixing zone between two guide plates.

Fig. 11
Turbulent heat transfer (a) and temperature gradient (b) in a direction perpendicular to the mainstream direction
Fig. 11
Turbulent heat transfer (a) and temperature gradient (b) in a direction perpendicular to the mainstream direction
Close modal

To sum up, since the streamwise vortex and the fluctuating velocity intensity enhance the mixing process of the hot air and the cold air, and the maximum intensity value of those always appeared in the area behind the two guide plates, it meant that installing ribbed airflow control doors in the flow path of HVAC unit which perform the same function as the guide plates was an effective method to promote mixing.

4 Determination of Turbulent Prandtl Number

4.1 Turbulent Prandtl Number.

The Prt is a classical approach for the heat transfer problem in turbulence, which is defined as the ratio between the momentum eddy diffusivity εm and the heat transfer eddy diffusivity εh, which can be written as follows:
(4)
where U is the magnitude of time averaged velocity, T is the time averaged temperature, the turbulent fluctuating terms uv¯ and vT¯ are the apparent turbulent shear stress and the apparent turbulent heat flux, respectively. Thus, the velocity field, the temperature field, and the turbulent fluctuating terms are known, the Prt can be evaluated. However, if the Prt value in the mixing zone is calculated using Eq. (4), there are concerns that adversely affect the calculation of temperature gradients and turbulent heat transfer term due to the insufficient spatial and temporal measurement resolution in the temperature field, and so for compensation of this defect, data assimilation is performed based on these experiments that simulate the mixing field inside the simple HVAC model and the Prt distribution that fits the experimental data is predicted. Data assimilation is a cross-disciplinary science to synergize the numerical simulation and the observational data, using statistical methods and applied mathematics. In recent decades, there is an increasing research on improving the accuracy of the CFD simulation using data assimilation, e.g., Kato et al. [19] simulated back step flow accurately by assimilating the turbulence model parameters in RANS-based CFD with the measurement data, Farhat et al. [20] applied a continuous data assimilation scheme to numerical simulation of Rayleigh Benard convection at infinite or large Prandtl numbers using only the temperature field as observables, Shimoyama et al. [21] applied the data assimilation to the porous media model and improved the accuracy of the RANS-based thermo-fluid dynamics simulation for an HVAC heat exchanger, etc.

4.2 CFD Simulation.

The CFD simulation model of the simple HVAC model used for the data assimilation reproduces the mixing field of the experiment in the previous section, and a polyhedral mesh with 927,594 cells is generated in the simulation domain as shown in Fig. 12. Regarding the error or numerical uncertainty in CFD simulation due to the mesh [22], we do not address in the paper and leave for future work since this study is only to focus on the effect of the Prt in temperature field simulation. According to our previous experience, the standard size of mesh is 5.0 mm and the mesh density is 1, the number of prism layers and the prism layer thickness are set to 2 and 33.3% at the wall of simulation model. The governing equations are the Reynolds averaged continuous equation, the Navier–Stokes equation, and the energy equation. The analysis assumes a three-dimensional incompressible steady flow [23] due to the Mach number which is less than 0.05 (the maximum velocity is 15 m/s). Here, the energy equation of the RANS model is generally expressed by the following equation:
(5)
where the subscript j is an index that identifies the three directions of the Cartesian coordinate system; x is the Cartesian coordinates; u¯j and T¯ are Reynolds averaged velocity and temperature, respectively; λ is the thermal conductivity; and υt is the eddy viscosity.
Fig. 12
CFD simulation model of simple HVAC
Fig. 12
CFD simulation model of simple HVAC
Close modal

The Prt is the model parameter for data assimilation estimation. From Eq. (5), the heat conductivity of the air changes with respect to the variation of the Prt, so the value of the Prt is a parameter that controls the heat transfer within the mixing field. Normally, it is often assumed that the value of the Prt is about 0.9 in the case of air, but the temperature field of simulation does not match the experiment. Then data assimilation modifies the value of Prt at each location so that the simulation results of the mixing field match the observed data in the experiment.

In the study, CFD simulation is carried out by a commercial CFD software Siemens star-ccm+ 13.06.012. Regarding the simulation model, the traditional setting for previously validated CFD simulation is employed as follows, the semi-implicit method for pressure-linked equation is used to solve the governing equations discretized by a finite volume method, and the quadratic precision upwind method using the gradient by the cell-based minimum square method is applied to each term of the Navier–Stokes equation, the realizable kε turbulence model is chosen here. As the boundary conditions, air velocity and temperature are fixed at the inlet, which are taken from the experimental conditions in Table 1, and the atmospheric pressure and temperature are imposed at the outlet, the turbulence intensity and the turbulence length scale are set to 1% and 10 mm at the inlet and outlet, and the two layer all wall y+ model is adopted on the wall of the HVAC model with no-slip condition and a constant thermal resistance.

4.3 Data Assimilation Techniques.

We use the ensemble Kalman filter (EnKF) [24] for data assimilation, which is a typical method of sequential data assimilation. In a normal Kalman filter, the time evolution of the covariance matrix is explicitly calculated, but in the extremely multivariate problems targeted by data assimilation methods, this covariance matrix becomes extremely large. The EnKF significantly reduces the computational cost of the Kalman filter by approximately representing the covariance matrix using large number of ensembles. Figure 13 shows the flowchart of the present EnKF for CFD simulation. The state vector x consists of the flow quantities (velocity component u, v; temperature T; and the Prt) calculated at Nflow mesh nodes in the CFD model of the simple HVAC (currently, Nflow = 927,594) as
(6)
and the observation vector y is a vector with 60 elements consisting of the x-direction average velocity u, the y-direction average velocity v, and the average temperature T at 60 measured points in the mixing zone as shown in Fig. 3:
Fig. 13
Flowchart of the EnKF for CFD simulation
Fig. 13
Flowchart of the EnKF for CFD simulation
Close modal
(7)

First in the flowchart, the method of generating the initial ensemble x0|0(i) (i = 1, 2, …, Nensemble = 927,594) is arbitrary, but here it was generated based on the Prt function of Ito et al. [4]. First, a convergent solution of the flow field is obtained using the Prt function under predetermined test conditions. The x-direction velocity u, y-direction velocity v, temperature T, and the Prt obtained are stored in the initial ensemble x0|0(i). Finally, an initial ensemble with a different Prt distribution is generated by adding random numbers to the Prt part of each ensemble member. Since each ensemble has smooth Prt distribution and a trend appears in the model parameter distribution, this method has an advantage that it is easier to construct the Prt function from the estimated Prt distribution, compared to an initial ensemble that was generated by simply throwing random numbers and adding them to a reference value.

Starting with n = 0, CFD is conducted for the i-th ensemble member xn|n(i) to predict the state vector at the next assimilation step, xn+1|n(i), and the corresponding observation vector, yn+1(i), which is linearly obtained through a mapping matrix H as
(8)
Next, EnKF generates a new ensemble, xn+1|n+1(i), as
(9)
Kn+1 is the Kalman gain given by
(10)
where V^n+1|n is the ensemble covariance matrix given by
(11a)
(11b)

In addition, w is the observation error vector, which is currently set as w ∼ Norm (0, R) with R = 0.1I (I is the identity matrix). The current value of the observation error is tentatively employed for successful data assimilation.

This process is repeated (n = n + 1) until the following cost function J is converged (hopefully, the predicted y approaches the actual measurement data in the simple HVAC unit, yactual, for one experimental condition among the 60 points)
(12)
where the subscript i is an index that identifies the observation point, the one with the subscript cal is the calculated value, and the one with exp is the measured value. This is the sum of squares of the differences at the measurement points in the mixing field. It is generally considered that the cost function value decreases with each step of data assimilation, and by calculating this at each step of data assimilation and looking at the history, the degree of assimilation with the observation of the calculation can be confirmed.

4.4 Results of the Prt.

Data assimilation is carried out under each experimental condition, and the Prt function [4] is used as a reference value for generating the initial ensemble; first obtain the convergence solution of the flow field by the reference value under the predetermined experimental conditions, since the Prt is a function of the flow field, and it is a smooth distribution at the end of the fluid analysis. Here, we create one ensemble member by adding one random number from the uniform distribution to the entire distribution.

The target of this data assimilation is to determine only the Prt, since the Prt only affects the temperature field and does not change the velocity field, and we will focus on the temperature field from now on. Figure 14 shows the calculation results of temperature using the Prt of the data assimilation estimation at each observation point under the conditions of test 03 and test 04. The blue diamond points are the measured values, the short vertical blue lines represent 95% confidence interval error bars, the green square points are the calculation results using the Prt estimated for data assimilation, and the red triangle points are the calculation results using the conventional function of the Prt by Ito et al [4]. The average temperature absolute errors for each measured temperature are 4.59 °C and 2.55 °C for data assimilation and 7.34 °C and 6.77 °C for the conventional Prt function, and this reduction in errors is confirmed under other experimental conditions as well, −1.82 °C for test 01, −2.88 °C for test 02, −0.47 °C for test 05, and −1.13 °C for test 06; it indicates that the prediction accuracy of the temperature field is improved by correcting the Prt for data assimilation. Here note that the deviation between the experimental data and the calculation result obtained through data assimilation increases with the numbered measurement points in the case of test 03 but such scenario is not there in the case of test 04. This is because that the high-numbered measurement points are installed close to the exit of the simple HVAC model, where the air flow is more unstable and turbulent in the case of test 03 due to high air velocity. Since the velocity field in the turbulent region is hard to resolve without considering the more precise turbulence model, the deviation of velocity similarly becomes larger with the numbered measurement points as shown in Fig. 15.

Fig. 14
Comparison of the calculation results and experimental data for temperature in the case of (a) test 03 and (b) test 04
Fig. 14
Comparison of the calculation results and experimental data for temperature in the case of (a) test 03 and (b) test 04
Close modal
Fig. 15
Comparisons of the calculation results and experimental data for velocity component in the case of (a) test 03 and (b) test 04
Fig. 15
Comparisons of the calculation results and experimental data for velocity component in the case of (a) test 03 and (b) test 04
Close modal

This is an issue to be addressed in future work. It indicates that the data assimilation leads to the converged CFD solution, which is consistent with the actual measurements of temperature as shown in Fig. 14. Regarding velocity shown in Fig. 15, on the other hand, there is still a large deviation even by the data assimilation in the case of high velocity air flow, and the same trend is observed under other test conditions. This may be because the discrepancy mainly comes from the turbulent model itself, and the current used turbulent model is simple and not suitable for real turbulent flow. For further improvement in CFD simulation accuracy assisted by the data assimilation, therefore, we are thinking of a possibility to optimize parameters of the turbulent model with data assimilation.

Next, Fig. 16 plots the variation of the Prt with air velocity under all test conditions. The triangles indicate the calculation results of Eq. (4), and several unrealistic data are excluded due to the lack of the measured data; the circles represent the results from data assimilation. It shows that the Prt is reducing with decreasing mainstream velocity, and when the mainstream velocity exceeds a certain value (uc = 5.0 m/s), the Prt becomes larger when velocity increases. In comparison with data assimilation results and calculation results of Eq. (4), it seems that the data assimilation results have a clear relationship with the flow velocity, and the Prt can be easily functionalized using the data assimilation result.

Fig. 16
Comparison of the Prt number between data assimilation results and the calculation results of Eq. (5).
Fig. 16
Comparison of the Prt number between data assimilation results and the calculation results of Eq. (5).
Close modal

4.5 Functionalization of the Prt.

Functionalization is a convenient way to apply the Prt data assimilation results to thermal fluid simulations of actual HVAC units, and it is generally thought of as an exponential function of Reynolds number Re and molecular Prandtl number Pr such as Prta×Reb×Prc. However, in this study, according to the variation characteristics of the Prt with respect to velocity (see Fig. 16), two functions for low velocity (U<uc) and high velocity (Uuc) regions are proposed, and then the key variables for the function of Prt are determined from several physical quantities by sensitivity analysis on a commercial optimization analysis software (modefrontier); here these physical quantities include magnitude of velocity U, temperature T, velocity gradient dU(U/x)2+(U/y)2, and temperature gradient dT=(T/x)2+(T/y)2. Figure 17 shows the results of sensitivity analysis, and it shows that two important variables are found to be U and T dominated in the low velocity region and only one key variable U dominates in the high velocity region. However, the influence of second sensitivity variable dT in the high velocity region is also considered in the study.

Fig. 17
The sensitivity analysis results of the key variables for the function of Prt
Fig. 17
The sensitivity analysis results of the key variables for the function of Prt
Close modal

Hereafter, for the purpose of distinguishing, the Prt function value and the Prt data assimilation results are denoted as Prt_Eq and Prt_DA, respectively.

According to our previous experience and intuition, the form of the Prt function was proposed as follows:
(13)
where these dimensionless variables are defined using reference parameters in Table 1, such as
(14a)
(14b)
Taking the logarithm of both sides of Eq. (13) obtains
(15)
where alorh*lnalorh. The coefficients (alorh*,blorh,clorh,dlorh) are determined from data assimilation results using multidimensional analysis, and the optimal square difference between the function calculated value Prt_Eq(i) and the data assimilation results Prt_DA(i) is represented as
(16)
where Nl and Nh are the measurement points of Prt_DA in the low velocity region and the high velocity region, respectively.
The ϕ is minimal when ϕ/alorh*=ϕ/blorh=ϕ/clorh=0; here each one is calculated in the case of low velocity as follows:
(17a)
(17b)
(17c)
Converting to determinant, we get
(18)
Then solving Eq. (18) gives al(al*), bl, and cl. Similarly for the case of high velocity, finally, the proposed functions Prt_Eq are
(19)

Figure 18 plots the variations of Prt_Eq and Prt_DA with respect to U, and the coefficient of determination R2 has a high value of 0.93 as shown in Fig. 19; it seems that the Prt function could roughly reproduce the data assimilation results. However, an issue is that the relative error is large in the small value region, it is conceivable that the amount of experimental data was insufficient, or that the current proposed Prt function is still not the optimal forms. Regarding improvement of the Prt function, there are two ways which are increasing experimental data under small velocity conditions and establishing more appropriate functional forms using optimization analysis. Anyway, if enough experimental data can be collected in the future, the machine learning model might be a final way instead of the Prt function.

Fig. 18
The variation of the number with air velocity
Fig. 18
The variation of the number with air velocity
Close modal
Fig. 19
The coefficient of determination R2
Fig. 19
The coefficient of determination R2
Close modal

5 Conclusions

In this study, data assimilation was performed based on the experiments that simulate the mixing field by hot and cold air inside a simple HVAC model, and the turbulent Prandtl number distribution that fits the experimental data were predicted. It was confirmed that the accuracy of the temperature field prediction is improved under all experimental conditions compared to the results using the conventional turbulent Prandtl number function by Ito et al. [4], as the average deviation of temperature between the experiment and simulation reduced by 2.2 °C and was less than 3.5 °C. It was demonstrated that the data assimilation is most effective for CFD simulation to simulate the temperature field consistent with the actual measurements. Consequently, the data assimilation is promising to enhance high fidelity as well as low cost in real\minus world product design and development. In addition, in order to apply the turbulent Prandtl number determined by data assimilation to CFD simulations of actual HVAC unit, two new turbulent Prandtl number functions are proposed, which are constructed from a power function of the dimensionless velocity U* and an exponential function of the dimensionless temperature T* or temperature gradient dT*, such as 0.016×U*0.187×e1.498×T* for the low velocity region and 0.040×U*0.274×e0.006×dT* for the high velocity region.

Conflict of Interest

There are no conflicts of interest.

Data Availability Statement

The datasets generated and supporting the findings of this article are obtainable from the corresponding author upon reasonable request.

Nomenclature

d =

distance between thermocouples

h =

heat transfer coefficient in the Kramers equation

w =

observation error vector of data assimilation

J =

cost function of data assimilation

T¯ =

Reynolds averaged temperature

U =

magnitude of velocity

H =

mapping matrix in data assimilation

I =

identity matrix of data assimilation

cw =

heat capacity of wire in the Kramers equation

kf =

thermal conductivity of fluid

rw =

radius of wire in the Kramers equation

uc =

certain value of velocity

u¯j =

Reynolds averaged velocity component

x^n+1|n =

ensemble state vector of data assimilation

yactual =

vector of actual measurement data

Nflow =

mesh nodes in the CFD model

Nensemble =

total ensemble member

Nl =

number of measurement points in the low velocity region

Nh =

number of measurement points in the high velocity region

Tf =

temperature of fluid in the Kramers equation

Th =

high temperature at inlet of a simple HVAC model

Tl =

low temperature at inlet of a simple HVAC model

Tw =

temperature of sensor in the Kramers equation

Uh =

high velocity at inlet of a simple HVAC model

Ul =

low velocity at inlet of a simple HVAC model

Kn+1 =

Kalman gain of data assimilation

V^n+1|n =

ensemble covariance matrix of data assimilation

T* =

dimensionless temperature of the Prt functions

U* =

dimensionless velocity of Prt functions

xn|n(i) =

state vector of data assimilation

yn+1(i) =

observation vector of data assimilation

alorh,alorh*,blorh,clorh =

coefficients of Prt functions

dT =

temperature gradient

dT* =

dimensionless gradient temperature of Prt functions

dU =

velocity gradient

uv¯ =

Reynolds shear component (turbulent shear stress)

vT¯ =

turbulent heat flux

v*T¯ =

turbulent heat flux at a local coordinate system

(xi,yi) =

Cartesian coordinate system

(xi*,yi*) =

local coordinate system

Pr =

molecular Prandtl number

Prt =

turbulent Prandtl number

Prt_DA =

Prt data assimilation results

Prt_Eq =

Prt function value

Prf =

Prandlt number in the Kramers equation

Re =

Reynolds number

Ref =

Reynolds number in the Kramers equation

εh =

heat transfer eddy diffusivity

εm =

momentum eddy diffusivity

λ =

thermal conductivity

ρw =

linear density of wire in the Kramers equation

τ =

thermal time constant in the Kramers equation

υt =

eddy viscosity

ϕ =

optimal square difference

ωz =

vorticity

Subscripts

cal =

calculated value

exp =

measured value

n|n =

current step|next step of assimilation

Superscript

(i) =

the i-th ensemble member

References

1.
Fujisawa
,
N.
,
1999
, “
Air Flow Measurement by Image Analysis and Numerical Calculation of Flow Fields in Heater Unit for Automobiles
,”
Trans. JSME B (in Japanese)
,
65
(
629
), pp.
48
53
.
2.
Misuno
,
Y.
,
Asano
,
H.
,
Tada
,
K.
,
Hirota
,
M.
,
Nakayama
,
H.
, and
Hirayama
,
S.
,
2005
, “
Turbulent Mixing of Hot and Cold Air in HVAC Unit for Automobile
,”
Trans. JSME B (in Japanese)
,
71
(
701
), pp.
215
222
.
3.
Tan
,
L. B.
, and
Yuan
,
Y. J.
,
2022
, “
Computational Fluid Dynamics Simulation and Performance Optimization of an Electrical Vehicle Air-Conditioning System
,”
Alexandria Eng. J.
,
61
(
1
), pp.
315
328
.
4.
Ito
,
K.
,
Liu
,
J.
, and
Onodera
,
J.
,
2017
, “
Investigation of CFD Temperature Prediction for HVAC Units
,”
Proceedings of the 2017 JSAE Annual Spring Congress (in Japanese)
,
Yokohama, Japan
, pp.
691
694
.
5.
Burazer
,
J.
,
Novkovic
,
Ð.
,
Cocic
,
A.
,
Bukurov
,
M.
, and
Lecic
,
M.
,
2022
, “
Numerical Study of the L/D Ratio and Turbulent Prandtl Number Effect on Energy Separation in a Counter Flow Vortex Tube
,”
J. Appl. Fluid Mech.
,
15
(
5
), pp.
1503
1511
.
6.
Blom
,
J.
,
1970
, “
Experimental Determination of the Turbulent Prandtl Number in a Developing Temperature Boundary Layer
,”
4th International Heat Transfer Conference
,
Paris—Versailles, II
, p.
1
, Paper FC2.2.
7.
Reynolds
,
A. J.
,
1975
, “
The Prediction of Turbulent Prandtl and Schmidt Numbers
,”
Int. J. Heat Mass Transfer
,
18
(
9
), pp.
1055
1069
.
8.
Browne
,
L. W. B.
, and
Antonia
,
R. A.
,
1983
, “
Measurements of Turbulent Prandtl Number in a Plane Jet
,”
ASME J. Heat Transfer
,
105
(
3
), pp.
663
665
.
9.
Kays
,
W. M.
,
1994
, “
Turbulent Prandtl Number. Where Are We?
ASME J. Heat Transfer
,
116
(
2
), pp.
284
295
.
10.
Morris
,
G. K.
,
Garimella
,
S. V.
, and
Amano
,
R. S.
,
1996
, “
Prediction of Jet Impingement Heat Transfer Using a Hybrid Wall Treatment With Different Turbulent Prandtl Number Functions
,”
ASME J. Heat Transfer
,
118
(
3
), pp.
562
569
.
11.
Churchill
,
S. W.
,
2002
, “
A Reinterpretation of the Turbulent Prandtl Number
,”
Ind. Eng. Chem. Res.
,
41
(
25
), pp.
6393
6401
.
12.
Redjem-Saad
,
L.
,
Ould-Rouiss
,
M.
, and
Lauriat
,
G.
,
2007
, “
Direct Numerical Simulation of Turbulent Heat Transfer in Pipe Flows: Effect of Prandtl Number
,”
Int. J. Heat Fluid Flow
,
28
(
5
), pp.
847
861
.
13.
Venayagamoorthy
,
S. K.
, and
Stretch
,
D. D.
,
2010
, “
On the Turbulent Prandtl Number in Homogeneous Stably Stratified Turbulence
,”
J. Fluid Mech.
,
644
, pp.
359
369
.
14.
Li
,
D.
,
2019
, “
Turbulent Prandtl Number in the Atmospheric Boundary Layer-Where Are We Now?
Atmos. Res.
,
216
, pp.
86
105
.
15.
Basu
,
S.
, and
Holtslag
,
A. A. M.
,
2021
, “
Turbulent Prandtl Number and Characteristic Length Scales in Stably Stratified Flows: Steady-State Analytical Solutions
,”
Environ. Fluid Mech.
,
21
, pp.
1273
1302
.
16.
Xu
,
X.
,
Ooi
,
A.
, and
Richard
,
D. S.
,
2021
, “
Data-Driven Algebraic Models of the Turbulent Prandtl Number for Buoyancy-Affected Flow Near a Vertical Surface
,”
Int. J. Heat Mass Transfer
,
179
, p.
121737
.
17.
The Visualization Society of Japan
,
2002
,
Handbook of Particle Image Velocimetry (in Japanese)
, Morikita Publ., Tokyo.
18.
Khine
,
S. M.
,
Houra
,
T.
, and
Tagawa
,
M.
,
2013
, “
Theoretical Analysis and Its Experimental Validation of Heat-Conduction Error of Temperature Sensors in a Flow Field With Mean-Temperature Gradient
,”
Therm. Sci. Eng. (in Japanese)
,
21
(
3
), pp.
73
81
. doi.org/10.11368/tse.21.73
19.
Kato
,
H.
,
Ishiko
,
K.
, and
Yoshizawa
,
A.
,
2016
, “
Optimization of Parameter Values in the Turbulence Model Aided by Data Assimilation
,”
AIAA J.
,
54
(
5
), pp.
1512
1523
.
20.
Farhat
,
A.
,
Glatt-Holtz
,
N. E.
,
Martinez
,
V. R.
,
McQuarrie
,
S. A.
, and
Whitehead
,
J. P.
,
2020
, “
Data Assimilation in Large Prandtl Rayleigh-Bénard Convection From Thermal Measurements
,”
SIAM J. Appl. Dyn. Syst.
,
19
(
1
), pp.
510
540
.
21.
Shimoyama
,
K.
,
Sato
,
Y.
,
Onodera
,
J.
, and
Liu
,
J.
,
2021
, “
Measurement-Based Strategies for High-Fidelity Thermo-fluid Dynamics Simulation of an Automotive Heat Exchanger
,”
J. Fluid Sci. Technol.
,
16
(
1
), p.
JFST0006
.
22.
Celik
,
I. B.
,
Ghia
,
U.
,
Roache
,
P. J.
,
Freitas
,
C. J.
,
Coleman
,
H.
, and
Raad
,
P. E.
,
2008
, “
Procedure for Estimation and Reporting of Uncertainty Due to Discretization in CFD Applications
,”
ASME J. Fluids Eng.
,
130
(
7
), p.
078001
. doi.org/10.1115/1.2960953
23.
Anderson
,
J. D.
,
1995
,
Computational Fluid Dynamics: The Basics With Applications
,
Mc Graw-Hill
,
New York
.
24.
Evensen
,
G.
,
2003
, “
The Ensemble Kalman Filter: Theoretical Formulation and Practical Implementation
,”
Ocean Dyn.
,
53
(
4
), pp.
343
367
.