Sunteți pe pagina 1din 10

COMPARISON OF THREE k-e TURBULENCE MODELS FOR PREDICTING VENTILATION AIR JETS

Q. Liu, S. J. Hoff, G. M. Maxwell, D. S. Bundy


ABSTRACT. Published data on unconfined plane-wall jets and plane-free jets were reviewed and used to assess the accuracy of numerical simulations, A plane-free jet was numerically simulated using the standard k-e model and four nonuniform grid patterns (70 x32,100 x52,120 x 60, and 120 x 74), The solution for a plane-free jet with adequate grid resolution was in good agreement with the published data, A plane-wall jet was numerically simulated usingfivedifferent grids (70 x 32, 100 x 52, 120 x 60, 120 x 74, and 120 x 92) and three k-e turbulence models (the standard k-e model. Lam and Bremhorst low Reynolds number model, and Lam and Bremhorst low Reynolds number model with wall functions). The simulations predicted velocity decay and velocity profile well, but overpredicted the jet spread and entrainment ratio by 20 to 40%, indicating the need for a better turbulence model for wall jet predictions. Keywords. Ventilation, Numerical, Grid pattern.

et-ventilated structures are widely used in both agricultural and nonagricultural systems. Numerous research has been done on ventilation air flows involving jets as well as on jets themselves. Because jets have been studied extensively, they provide good test cases to estimate the effects of using different turbulence models and grids. Turbulence modeling is an important part of numerical simulation. Currently, most of the ventilation simulations use one of three k-e models (k is turbulent kinetic energy, e is the dissipation rate of turbulent kinetic energy): the standard k-e model (Launder and Spalding, 1974), Lam and Bremhorst (1981) low Reynolds number model, or the Lam and Bremhorst low Reynolds number with wall functions (Chen et al., 1990). Aldiough most of the studies reported agreement with the corresponding experimental results using the three models, detailed comparison of the performances of the three models is not available in a ventilation setting. Figure 1 is a sketch of a plane-free jet and a plane-wall jet. A plane-free jet is a two-dimensional jet fliat is not attached to any solid surface. If a two-dimensional jet is attached to a solid surface, the jet is a plane-wall jet. For ventilation applications, the jet characteristics of interest include the velocity profile, velocity decay (or peak velocity), jet spread, and entrainment ratio. The velocity

profile is the velocity profile at a given x location (fig. 1). The velocity decay is the peak velocity u^ as a function of X. The jet spread is the location of yo.5 as a function of x. The entrainment ratio is the flow rate at location x divided by the inlet flow rate, Qx/QoThe objective of this study was to compare the performances of three k-e turbulence models in predicting the jet characteristics of interest to ventilation. Available experimental data were used to assess the performance of numerical simulation and the impacts of different grids and turbulence models. Since inadequate grid resolution can be a major source of error (Thangam and Speziale, 1992), five different grid patterns were used to ensure adequate grid resolution and to evaluate the impact of using different grids in the numerical simulation of jets. Published data on unconfined isothermal air jets were summarized in terms of velocity profile, jet spread, velocity decay, and entrainment ratio for both plane-free and plane-wall jets. The three models used in this study were selected from available turbulence models (Liu et al..

Article was submitted for publication in April 1995; reviewed and approved for publication by the Structures and Environment Div. of ASAE in September 1995. Journal Paper No. J-16264 of Iowa Agriculture and Home Economics Experiment Station, Ames, Iowa. Project No. 3140. The authors are Qianbao Liu, ASAE Member Engineer, Assistant Professor, Dept. of Agricultural Engineering, University of Puerto Rico, Mayaguez, Puerto Rico; Steven J. Hoff, ASAE Member Engineer, Associate Professor, Dwaine S. Bundy, ASAE Member Engineer, Professor, Agriculmral and Biosystems Engineering Dept., and Gregory M. Maxwell, Associate Professor, Mechanical Engineering Dept., Iowa State University, Ames. Corresponding author: Steven J. Hoff, 105 Davidson Hall, Agricultural and Biosystems Engineering, Iowa State University, Ames, lA 50011; telephone: 515-294-6180; e-mail: <hoffer@abel .ae.iastate.edu>.

..
b

i
M).9

AUn/2 ^Un

!
Figure 1-A plane-free jet and a plane-wall j e t

Transactions of the ASAE VOL. 39(2):689-698 1996 American Society of Agriculmral Engineers 0001-2351 / 96 / 3902-0689

689

1995a) because they were conmionly used in ventilation simulations. The jets studied were a plane-free jet and a plane-wall jet. The jets were isothermal and unconfined (except confinement due to a soUd wall that form a wall jet for the plane-wall jet). The inlet height b (half of the inlet height for the plane-free jet) was 0.05 m and the inlet velocity was 6 m/s and was assumed uniform across the inlet. The Reynolds number based on inlet height and mean velocity was 2.0 X 10^.

LITERATURE REVIEW
EXPERIMENTAL AND ANALYTICAL STUDIES ON PLANE JETS

Unconfined air jets can be divided into four zones (ASHRAE, 1989)a short zone in which the maximum velocity of the jet remains practically unchanged (zone 1), a transition zone (zone 2), a zone of fully established turbulent flow, which is also referred to as the self-similar zone (zone 3), and a terminal zone (zone 4). Zone 3 is the longest and the most important in engineering applications among the four zones. The results reviewed in tiiis study are for zone 3 only. However, the division line between zones 3 and 4 may need clarification. The suggested division line is at x/b of about 100 (ASHRAE, 1989). Many experiments measured jets at x/b over 100 (Launder and Rodi, 1981; Narasimha et al., 1973). The results showed that zone 3 behavior continued at x/b > 100 and as high as x/b =1000. Plane jets have been studied in terms of velocity profile, jet spread, turbulence properties, and shear stress (Koestel et al., 1950; Tuve, 1953; Sigalla, 1958; Schwarz and Cosart, 1961; Myers et al., 1963; Black et al., 1969; Wilson et al., 1970; Walker, 1977; Wygnanski et al., 1992). The velocity decay of a wall jet was found to be equivalent to that of a free jet having twice the inlet width of the wall jet (Tuve, 1953; Walker, 1977; ASHRAE, 1989). The same was true for the velocity profile of a wall jet except for the

region close to the wall (fig. 3b and fig. 5b), although the velocity expressions were different (table 1.). Tuve (1953) found that the wall-jet spread was slightly less than that of the corresponding free jet. Other studies quantified the difference as about 30% less than the corresponding free jet (Launder and Rodi, 1981; Rajaratnam, 1976). Because of the difficulty in measuring flow rate, the entrainment ratio was usually calculated using the velocity profile, jet spread, and velocity decay (Rajaratnam, 1976). The entrainment ratio of the plane-wall jet was also about 30% less than the corresponding free jet due to the reduced jet spread. However, the jet spread and entrainment ratio calculations for wall jets were assumed to be equivalent to the corresponding free jet (ASHRAE, 1989; Albright, 1990). Studies on both plane-free jets and plane-wall jets were reviewed and summarized by Rajaratnam (1976) and Abramovich (1963). Launder and Rodi (1981) reviewed experimental studies on plane-wall jets. Fifteen studies on plane-wall jets were compared for velocity decay, jet spread, turbulence intensity, momentum loss along the jet, and other properties. Using experimental data from 17 studies, Narasimha et al. (1973) showed that inlet momentum flux JQ (JQ = ^o^b) rather than both UQ and b could be used to determine "most (if not all) mean flow parameters" in the fully developed region (x/b > 30). In their experimental study, Wygnanski et al. (1992) reaffirmed the Narasimha et al. (1973) finding on inlet momentum flux. Table 1 is a list of results reported in the literature and the results were used to assess the accuracy of the numerical simulations. The results cited here are for the self-similar portion of the jet (zone 3) only. Refer to nomenclature for all variables used in the table.
NUMERICAL SIMULATIONS

A number of numerical simulations on jets have been conducted. Setrak (1988) used the standard k-e model (Launder and Spalding, 1974) and SIMPLE (Patankar,

Table 1. Summary of jet characteristics Equation and Number Velocity profile (free jets) -y--exp(-0.693Ti2)
m

Sources (1) Rajaratnam (1976), ASHRAE (1989)

Velocity profile (wall jets)

J L . I.48TI1/7[1 - erf(0.68n)]
m

(2)

Rajaratnam (1976)

Jet spread (free jets) Jet spread (wall jets) Velocity decay (free jets)

yo.5 = 0.1x yo.5=0.073(x+12b)

(3) (4) (5)

Awbi (1991), Rajaratnam (1976) Launder and Rodi (1981), Schwarz and Cosart (1961) Rajaratnam (1976)

"o Velocity decay (wall jets) Entrainment ratio (free jets)

V X Rajaratnam (1976), Walker (1977), ASHRAE (1989) (6) Walker (1977)

The same as equation 5

(7) 3 L 0.40A/^

Qo
Entrainment ratio (wall jets)

V b
(8) Equations 2, 4, and 5

3 L 0.28A/^+12

Qo

vb

690

TRANSACTIONS OF THE ASAE

1980) to simulate a wall jet and the effects of obstacles on the wall jet. An empirical entrainment ratio was embedded in the boundary condition at the free boundary of the wall jet. Simulation of a confined wall jet was part of the study done by Choi et al. (1988) using the standard k-8 model and a modified TEACH program. One constant in the standard k-e model, c^, was adjusted from 0.09 to 0.05 for wall jet simulation while 0.09 was used for other simulations. The predicted results were comparable with the experimental results. Grid patterns of 12 x 12, 22 x 22, 32 X 32, and 42 x 42 were tested. Rodi and Spalding (1970) and Ng and Spalding (1972) also simulated wall jets in an earlier search for better turbulence models. In a review of experimental and numerical studies on wall jets. Launder and Rodi (1983) concluded that the wall jet is a significantly more complex flow than a simply strained boundary layer flow for numerical simulation. Numerical studies of wall jets using three types of turbulence models were reviewedmodels using mixing length hypothesis, two-equation models including the standard k- model, and the Reynolds stress models. It was found that numerical simulations using the standard k-e model predicted a jet spread of about 30% greater than experimental results for a plane-wall jet in stagnant surroundings. They pointed out that this discrepancy was due to the fact that C^ was set as a constant in the model. By evaluating the standard k-e model using direct numerical simulation, Cazalbou and Bradshaw (1993) also showed that a number of constants (including C^) in the standard k-e model were different from region to region in wall-bounded flows (boundary layer flow and channel flow). Mixing length models needed to adjust the constants from flow to flow in wall jet prediction. Because of this problem, the mixing length model by itself was not popular in engineering applications (Nallasamy, 1987). Reynolds stress models were better than the two equation models and the models using mixing length hypothesis for predicting wall jets. The use of Reynolds stress models for engineering were limited because of the limited gains by using the models and the greater CPU time requirement (Bradshaw, 1987). The standard k-e model is widely used to predict ventilation air flows (Worley and Manbeck, 1995; Gan and Awbi, 1994; Knappmiller and Kirkpatrick, 1994; Maghirang and Manbeck, 1993; Chen and Jiang, 1992; Murakami et al., 1991; Choi et al., 1990; Fang and Grot, 1990; Awbi, 1989; Choi et al., 1988). However, two problems are commonly reported using the standard k-e model in engineering applications. One is that all the grid points have to be located in the fully turbulent region and wall functions are needed to bridge the solid surface and the first grid point adjacent to the solid surface (Launder and Spalding, 1974). Hoff (1990) found that his numerical simulation might not converge if all the grid points were not in a fully turbulent region. Choi et al. (1988) found that a fine grid of 42 x 42 did not converge but that coarser grids converged, possibly a result of the same problem. Another problem of the standard k-e model is the poor accuracy when predicting complex flows, especially flow with separation (White, 1991; Wilcox, 1993; Patel et al., 1991). Jet separation and reattachment are commonly encountered in ventilation air flows. Because of these problems, Hoff (1990) used a low Reynolds number model proposed by Lam and Bremhorst
VOL. 39(2):689-698

(1981). The Lam and Bremhorst low Reynolds number model compared favorably with a number of turbulence models (Liu et al., 1995a; Patel et al., 1984). Chen et al. (1990) used the Lam and Bremhorst low Reynolds number model with wall functions to predict water and air flows in a model and found that this combination was better than the standard k-e model. The combination of Lam Bremhorst model with wall functions were also used in Chen et al. (1992), Baker et al. (1994), and Baker and Kelso (1990). In more recent studies, Bergstrom (1994) used an algebraic stress model with a low Reynolds number model in a two region manner. Murakami et al. (1994) compared the performances of the standard k-e model, an algebraic model, and a Reynolds stress model and found that the standard k-e model was not as accurate as the other two models. As reviewed by Liu et al. (1995a), many new turbulence models have been proposed and some of them showed significant improvement over the standard k-e model. Most of the new models reviewed have not been tested for ventilation simulation. Most of the numerical simulations in ventilation applications use the SIMPLE(R) algorithm (Patankar, 1980). Detailed discussions of this algorithm and its application for ventilation simulation were given by Awbi (1991), Jones and Whittle (1992), and Hoff (1990). The importance of grid resolution was addressed by Baker and Kelso (1990) and Thangam and Speziale (1992). As pointed out by Baker and Kelso (1990), the problem of numerical diffusion may dominate the transport process without adequate grid resolution. Using a grid of 200 x 100, Thangam and Speziale (1992) showed that the reported errors in predicting flow over a back-facing step were mainly due to inadequate grid resolution.
NUMERICAL SOLUTION GOVERNING EQUATIONS

The governing equations for steady state flows can be written in the general form of (White, 1991; Patankar, 1980): ^fpUi<t>--r,,^|^)=S, (9)

where (|) = scalar variable FA gf^ = effective diffusion coefficient for (]) SA = source term for < [ ) The governing equations include continuity, momentum, turbulent kinetic energy, and dissipation for turbulent
Table 2. Summary of the partial differential equations Equation Continuity Momentum 1
Ui
^<|>,eff

0
^i + ^At

Turbulence energy Dissipation

k e

^i+[^lt/ak] \i+[li^/o^]

P(G-e) (pe/k)(CiG-C2e)

691

kinetic energy. The terms <[), F^^eff' ^^^ ^i^ ^^ summarized in table 2 for each equation. Variables used in table 2 are (Launder and Spalding, 1974): Ht-C^k2/e G'

Table 3. List of grids used Grid No. 1 Points inside inlet 1 Mesh* 70x32 X progression factor 1.08 y progression factor 1.1 Minimum y+t 111 1134 Maximum y+f Grid No. 2 5 100x52 1.1 1.1 23 77 Grid No. 3 8 120x60 1.08 1.1 15 48 Grid No. 4 15 120x74 1.08 1.1 7 26 Grid No. 5 30 120x92 1.09 1.1 0.8 9

\3y 3x1

U/

lay/

Number of points in x direction by the number of grids in y direction. For plane-wall jet only.

Ok - 1.0, Oe - 1.3, C^^ = 0.09f^, C, - 1.44f,, Cj = 1.92f2 For STD model (Launder and Spalding, 1974): f^.fi-f2-l For LB and LBW models (Lam and Bremhorst, 1981): f^= [l - e x p (- 0.0165Ry)]2/l + 2Q=5\ region where y*" > 11.5 (y*" calculated as k^^^ymin/^)- This combination allows the model to be used in low Reynolds number regions and does not require a dense grid in the solid surface region because wall functions are used. However, the performance of the LBW model near the solid surface region where y+ < 11.5 is questionable. Under this condition, the LBW model is the same as the LB model, however, the grid is not likely to be as dense as required by the LB model. All three models were used in this study for the planewall jet. For plane-free jets, only the STD model was used because the three models are essentially the same for free shear flows.
SELECTION OF GRIDS

f2=l-exp(-R^^

^ ve

ymin"" ^^ shortest distance to any solid surface


TURBULENCE MODELS

Three turbulence models were used: the standard k-e model (Launder and Spalding, 1974), referred to as the STD; the low Reynolds number model proposed by Lam and Bremhorst (1990), referred to as the LB; and the Lam and Bremhorst low Reynolds number model with wall functions (Chen et al., 1990), referred to as the LBW. The governing equations for each model are listed in table 2 with other equations. The STD model has been used successfully in numerous applications, especially for free shear flows. However, as reported in tiie literature, it requires all the grid points to be in the fully turbulent flow region and it is not accurate for complex flows. The LB model is intended to overcome the problems that the STD model has. It uses damping functions to extend its use to the low Reynolds number region and no wall functions are needed. The LB model requires a dense grid in the solid surface region and has relatively poor convergence behavior (Patel et al., 1984; Chen et al., 1990). The damping functions used also need improvement (Patel et al., 1984). The LBW model differs from the LB model in that wall functions are used in the solid surface
692

Five different grids were selected (table 3). The main difference between grids was the number of grid points in the jet region. Grid No. 1 was the coarsest grid which had one grid point inside the inlet. Grid 3 was the densest grid used for the STD model, which requires y+ > 11.5 for all the grid points. Grid 5 was the densest grid used with 30 grid points located inside the inlet. It had y+ values ranging from 0.8 to 9 for the grid points adjacent to the solid surface with all of them located in the region where y+ < 11.5. Compared with other low Reynolds number model studies using grid points located with y " * " values less than 0.2 (Wilcox, 1993; Chen and Patel, 1988), using grid 5 may not be dense enough for the LB model (Lam and Bremhorst, 1981). However, the grid used was much denser than grids commonly used for ventilation simulation. A denser grid was not attempted. The y+ values for grid points adjacent to the solid surface (for plane-wall jets only) are also shown in table 3. Table 4 lists the combinations of grids and turbulence models considered. Plane-wall jet simulations using grids 4 and 5 with the STD model did not converge. As shown in table 3, not all the grid points in grids 4 and 5 were in a
Table 4. Combination of grids and turbulence models used Grid No. 1 Grid No. 2 Grid No. 3 Grid No. 4 Grid No. 5

Plane-free Jet STD LBW LB


X
X X

Plane-wall Jet STD LBW LB


X X X X X X X X* X X X* X X

Convergence not achieved.

TRANSACTIONS OF THE ASAE

fully turbulent region. Grids 1 and 2 were not considered for the LB model because the LB model requires a dense grid in the solid surface region.
TREATMENT OF THE EXPONENTIAL TERMS

|B-

DOMAIN DF INTEREST

During the initial stage of this study, it was noticed that the LB and the LBW models, if used as proposed with the exponential terms, required about three times the CPU time that the STD model required to complete one iteration. The problem was traced to the exponential terms in the Lam and Bremhorst low Reynolds number model. To solve this problem, the exponential term e~^ was divided into five sections and each section was approximated by a polynomial. The division line between each section was drawn by observing the e~^ curve and minimizing the curvature in each section. Each polynomial was a least square fit of data points of the exponential function for the section. The polynomials are: e-^ /0.9962-0.9429x+0.3203x2 0.8146-0.5972X+0.1582x2-0.01471x3 0.2204-0.09677x+0.01447x2-7.325xlO-V 0.1009-0.03759x+0.004744x2-2.021xl0-V 0 (10)

1
jet>

wall boundary (wall jet> or centerllne of j e t ifree symmetry line exit

Figure 2-The computation domain. DETAILS OF THE NUMERICAL SCHEME

o<x< 1 l<x<3.9 3.9<x<5.5 5.5<x<8 x>8

This approximation has a relative error of < 2% in the range of 0 < x < 8 and an absolute error of < 0.0038 in the range of 0 < x < oo. Use of the approximation improved the speed of the LB and the LBW models to a rate comparable of the STD model.
BOUNDARY CONDITIONS

The computation domain is shown in figure 2. A computation domain of 250b in the x direction and 150b in the y direction was used in this study. The 250b was selected to ensure that the size of the domain was long enough to cover the suggested self-similar region of up to 100b (ASHRAE, 1989). The y direction size was arbitrary yet large enough that little impact was expected on the prediction. To minimize the complication of a solid surface, the left (except the inlet) and bottom boundary (AB and AE in fig. 2) were selected as symmetry lines. Uniform condition was assumed at the inlet for all the grids used to ensure comparability between grids. The e value at the inlet was calculated as 8 = (C^^^^^/^) / 0.4IL (Djilali et al., 1989; Gan and Awbi, 1994; Choi et al., 1988). L = 2b was used in the study. The boundary conditions are listed in table 5.

The SIMPLER (Patankar, 1980) method was used to numerically solve the partial differential equations. Central difference scheme was used to discretize the source terms. DEC 3000 workstations from Digital Equipment Corporation were used for the calculation. The convergence criterion was that the residual of the continuity equation should be < 2.0 x lO"-^. The flow rate at any given axial location x was needed for entrainment ratio calculation. The flow rate was calculated by integrating the velocity profile from the center of the jet (the wall in the case of wall jet) to the point where the air velocity component in the x direction was zero. It was found that the STD model was less likely to diverge. The guessed velocity field to start the iteration process may result in divergence using either the LBW or LB models, so the STD model solution was used as the initial guess for the LB and LBW models.

RESULTS AND DISCUSSION


The predicted results of velocity profile, velocity decay, jet spread, and entrainment ratio are presented. The parameters k, e, skin friction coefficient, etc. are not summarized here because these are not of primary interest for ventilation studies. Numerous discussions related to these parameters are available (Chen and Patel, 1988; Patel et al., 1984; Lam and Bremhorst, 1981; Wilcox, 1993). As mentioned earlier, uniform conditions were assumed across the inlet. The impact of that assumption on numerical simulation was not known. It was suggested that the inlet conditions might be important to achieve similitude (Liu et al., 1995b), however, the inlet condition

Ikble 5. Boundary conditions AB AE BC DE CD, plane-free jet CD, wall jet STD model and LBW model (y-- > 11.5) CD, wall jet, LB model and LBW model (y+ < 11.5) u = 0, dv/dx - 0, dk/Bx = 0, de/dx - 0 au/ay - 0, V = 0, dk/ay = 0, ae/ay - 0 u = uo, V - 0, k - 0.005uo2, e = 0.2k3/2 u - 0, Bw/dx = 0, dkldx = 0, dz/dx = 0, the momentum at DE should be less than or equal to the momentum at the adjacent upstream section*

au/ay = 0, V = 0, ak/ay = o, ae/ay = o


wall functions (Launder and Spalding, 1974; DjilaH et al, 1989; Chieng and Launder, 1980) u - 0, V - 0, k - 0, and e - 2vk/ymin2 (Chieng and Launder, 1980; Chen et al., 1990)

This condition was imposed to eliminate the possibility of unrealistic inflow at the exit. VOL. 39(2):689-698 693

was generally considered to have little impact on the selfsimilar portion (zone 3) of the jet (ASHRAE, 1989; Launder and Rodi, 1983; Narasimha et al., 1973).
PLANE-FREE JET

30

24

^
Eqn.(3) o Grid 4 (120x74) A Grid 3 (120x60) V Grid 2 (100x52) - Grid 1(70x32)

18

The plane-free jet was simulated using the STD model and four of the five grids as shown in table 3. The results are shown in figures 3 and 4. As shown in the figures, grids 2, 3, and 4 showed less than 5% difference between each other in the predicted results, indicating adequate resolution with grid 2. Grid 1, which was the coarsest grid with one grid point inside the inlet width, overpredicted the jet spread and entrainment ratio and underestimated the velocity decay. The results of the plane-free jet with adequate grid resolution (grids 2, 3, and 4) showed good agreement with published data. The velocity decay was virtually the same as equation 5. The jet spread prediction is slightly higher (< 10% higher) than the experimental result of equation 3. The entrainment ratio showed a slightly lower value than the experimental result (eq. 6) near the exit.
PLANE-WALL JET

,-"
12
y"

50

100 x/b

150

200

250

(a)
. V

W]
Eqn. (7) o Grid 4 (120x74) A Grid 3 (120x60) V Grid 2 (100x52) - Grid 1(70x32)

Jfi
/ ^

w
50 100 150 x/b 200 250

The plane-wall jet was simulated using the STD, LB, and LBW models with some or all of the five different grids (table 3). The STD model was used with grids 4 and 5, but convergence was not achieved. Both grids had grid points in the wall region with y+ < 11.5. Effects of Different Grids. The effects of different grids using the LBW model are shown in figures 5 and 6. The effects using the STD and the LB model were similar
1.0-

(b)

Figure 4-Resiilts of the plane-free jet simulation - (a) jet spread and (b) entrainment ratio.

0.8

Eqn. (5) o Grid 4 (120x74) A Grid 3 (120x60) ^ Grid 2 (100x52) - Grid 1(70x32)

Eqn. (5) o Grid 1(70x32) A Grid 2 (100x52) ^ Grid 3 (120x60) - Grid 4 (120x74) n Grid 5 (120x92)

50.6

0.4

0.2 0

1 50

i 100 x/b

1 150

" " i 200

250 250

(a)

(a)

Eqn.(l) o Grid 4 (120x74) A Grid 3 (120x60) ^ Grid 2 (100x52) - Grid 1(70x32)

Eqn. (2) o Grid 1(70x32) A Grid 2 (100x52) ^ Grid 3 (120x60) - Grid 4 (120x74) D Grid 5 (120x92)

y/yO.5

y/yO.5

(b)

(b)

Figure 3-Results of the plane-free jet simulation - (a) velocity decay and (b) velocity profile.
694

Figure 5-Results of the plane-wall jet simulation using the LBW model - (a) velocity decay and (b) velocity profile.
TRANSACTIONS OF THE ASAE

30 24 = 18

Eqn.(4)

J^?-;
^

.k,^ '

o Grid 1(70x32) A Grid 2 (100x52) V Grid 3 (120x60) Grid 4 (120x74) n Grid 5 (120x92)

6 0

^ ^^
100 x/b 150 200 250

50

(a)

Eqn.(8) o Grid 1(70x32) A Grid 2 (100x52) V Grid 3 (120x60) - Grid 4 (120x74) a Grid 5 (120x92)

250

(b) Figure 6--Results of the plane-wall jet simulation using the LBW model - (a) jet spread and (b) entrainment ratio.

difficult to judge what should be the adequate grid resolution. Comparing the results of grids 1, 2, and 3, all of which had y"^ > 11.5 and used wall functions, grid 2 showed larger overprediction of jet spread and entrainment than grids 1 and 3. Changing from grid 4 to grid 5 also changed the predicted results. Further investigation is needed to study if the same grid impact is true for jets with separation. Effects of Different Turbulence Models. The effects of different turbulence models using grid 3 are shown in figures 7 and 8. The results of using other grids were similar to grid 3. Figure 7a showed the velocity decay predictions using different turbulence models. The predicted results were all in good agreement with the experimental results. The STD model and LBW model gave identical results. The LB model prediction showed a slightly larger error than the other two models. The velocity profiles at x/b 150 are shown in figure 7b. The three models showed almost identical results and the results were in good agreement with the published data. However, the predicted profiles showed a peak velocity closer to the wall than what was observed in experiments. The jet spread predictions with different turbulence models are shown in figure 8a. All models significantly overpredicted the jet spread. The STD model and tiie LBW model gave predictions that were very close to each other and were about 30% larger than the experimental results. The LB model overpredicted the jet spread by about 40%. The 30% difference between the experimental results and the STD model prediction was in agreement with the conclusion of Launder and Rodi (1983). They concluded
1.0

to the LBW model. Figure 5a shows the velocity decay using different grids. As shown, all predictions were within 10% of the experimental results. Velocity profiles predicted are shown in figure 5b. The velocity profiles were at the location of x/b closest to 150 (depending on the grid, x/b ranged from 145 to 155). The velocity profiles showed little difference at different locations except the initial part of the jet (up to x/b = 40) and at the exit. The predictions were within 10% of the experimentally measured profile for most of the jet and within 30% at the outer region of the jet where the velocity was low. Predictions using grids 1 and 4 showed slighdy larger errors than the other grids. Figures 6a and 6b show the predicted jet spread and entrainment ratios. The jet spread predictions were 20 to 30% higher than the experimental results (fig. 6a). Grid 4 gave the best performance in terms of jet spread prediction. Grid 5 was the densest grid used but it did not have the best performance. The reason may be that all the grid points adjacent to the wall boundary for grid 5 have y+ values less than 11.5 (table 3) where LBW is effectively the same as the LB model. As shown later in the results, the LB model overpredicts the jet spread. The entrainment ratios were overpredicted by about 30%, with grids 1, 4, and 5 slightly better than the other two grids (fig. 6b). Grid 4 did not show a clear advantage in entrainment ratio prediction as it did for the jet spread prediction. The overpredictions of that grid in velocity profile and velocity decay eroded the advantage in jet spread. Based on the above analysis, no grid showed a clear advantage in terms of better agreement with the experimental results for the plane-wall jet, making it
VOL. 39(2):689-698

0.8o STD A LBW - LB Eqii.(5) 0.4-

10.6

0.250 100 x/b 150 200

< ^
250

(a)

(b) Figure 7-Results of the plane-wall jet simulation using grid 3 (a) velocity decay and (b) velocity profile.

695

30

CONCLUSIONS

o STD A LBW - LB Eqn.(4)

24 = 18 i

- "QS

s *

^l "C^

> - ^

6 0

M
DO "

^ ^
100 x/b 150 200 250

50

(a)

r^*"*" j^' ^ ^

- "A 6^

o STD A LBW LB Eqn.(8)

o"*

Y J L
/ _
50

100

150 (b)

200

250

Figure 8-Resiilts of the plane-wall jet simulation using grid 3 - (a) jet spread and (b) entrainment ratio.

that "the main cause of this discrepancy is the wall's damping of the normal velocity fluctuations". This damping is "not just in the immediate vicinity of the surface, but in the outer 'free-shear-layer' region beyond the velocity maximum". Similar conclusions were also drawn by Cazalbou and Bradshaw (1993). The LBW and LB model did not show any improvement over the STD model in terms of better agreement with the published data. The overprediction in jet spread resulted in the overprediction of the entrainment ratio (fig. 8b). The STD model and the LBW model showed almost identical results which were better than the LB model. Based on this analysis, it is concluded that the LB model is not as good as the other two models for wall jet prediction. This is especially true for jet spread prediction. The STD model and the LBW model gave comparable results. However, the STD model did not converge if grid points were not in the fully turbulent region (grids 4 and 5). The LBW model had an advantage in that it could be used for a region where the flow was not fully turbulent. The overprediction in jet spread and entrainment ratio may also exist in room air movement involving wall jets. The overprediction is likely to be smaller than that of an unconfined jet because the surroundings also affect jet expansion. A room of the same length will have a smaller overprediction with a lower room height because the jet's expansion is more likely to be restricted. Further studies on turbulence models are needed to improve the performance of jet spread prediction.

Published data on unconfined plane-free and planewall jets were summarized. It was found that the jet velocity decay, velocity profile, jet spread, and jet entrainment ratio were well established by experimental and analytical studies. Contrary to the common practice that an unconfined plane-wall jet can be treated as one half of the free jet, experimental results showed that the plane-wall jet has 30% less jet spread and entrainment ratio than the corresponding plane-free jet. For the plane-free jet, numerical simulation using the STD model showed that adequate grid resolution was reached with five grid points in the inlet for the jet studied. The solution was in good agreement with the experimental results. A grid that was too coarse (one grid point inside the inlet) overpredicted jet spread and underpredicted velocity decay for the plane-free jet. For the plane-wall jet, the LBW model was better than the other two models. The LB model gave the worst performance for the grids tested and the STD model may not converge if all the grid points are not in the fully turbulent region. All three turbulence models gave prediction of velocity decay and velocity profile close to the published data. But they all overpredicted the jet spread and entrainment ratio significantly (20 to 40%), indicating a need for better turbulence models to predict plane-wall jets. For the plane-wall jet, no one grid showed a clear advantage over the others among the five grids used. Error due to inadequate grid resolution may have offset other errors for the coarse grids. Because of the exponential terms used in the low Reynolds number model, the LB and LBW models took about three times as long to complete one iteration compared with the standard k-e model. Using a polynomial approximation to replace exponential terms significantly improved the speed of the low Reynolds number model used.

NOMENCLATURE
VARIABLES

jet inlet height for wall jet, half of the inlet height for plane-free jet Ci,C2, CH turbulence model constants error function: erf erf(x) = - |
-t^dt

G k Jo P Q Rk Rt

turbulence model functions turbulence generation specific turbulent kinetic energy momentum flux at the inlet mean pressure volumetric flow rate turbulence Reynolds number (k^/^y/x)) turbulence Reynolds number [k2y/(ve)]

696

TRANSACTIONS OF THE ASAE

U, V Um

y^ yo.5
ymin

x,y
< >

1 ^
V

Mt
^t'^e

P
11
SUBSCRIPTS

generalized source term velocity components jet velocity decay or peak velocity dimensionless distance from the solid surface location where u = u^^/l minimum distance to any solid surface dissipation rate of k coordinates generalized variable laminar dynamic viscosity laminar kinematic viscosity turbulent viscosity diffusion Prandtl numbers density diffusion coefficient ^ = y/yo.5 turbulent effective grid number at inlet X location

t off ij
0 X

REFERENCES
Abramovich, G. N. 1963. The Theory of Turbulent Jets. Cambridge, Mass.: The M.I.T. Press. Albright, L. D. 1990. Environment Control for Animals and Plants. St. Joseph, Mich.: ASAE. ASHRAE. 1989. ASHRAE Handbook (Fundamentals). Atlanta, Ga.: ASHRAE. Awbi, H. B. 1991. Ventilation of Buildings. London, England: E&FNSPON. . 1989. Application of computational fluid dynamics in room ventilation. Building and Environment 24:73-84. Baker, A. J. and R. M. Kelso. 1990. On validation of computational fluid dynamics procedures for room air motion prediction. ASHRAE Transactions 96(1): 760-774. Baker, A. J., P. T. Williams and R. M. Kelso. 1994. Development of a robust finite element CFD procedure for predicting indoor room air motion. Building and Environment 29(4): 261-273. Bergstrom, D. J. 1994. Numerical prediction of a turbulent fountain in a room. ASHRAE Trans. 100(2):669-676. Black, T. C , J. N. Walker and G. M. White. 1969. Plane non-isothermal air jets discharging along a smooth ceiling. ASAE Paper No. 69-450. St. Joseph, Mich.: ASAE. Bradshaw, P. 1987. Turbulent secondary flows. Ann. Rev. Fluid Mech. 19:53-74. Cazalbou, J. B. and P. Bradshaw. 1993. Turbulent transport in wall-bounded flows. Evaluation of model coefficients using direct numerical simulations. Physics of Fluids 5(12):3233-3239. Chen, H. C. and V. C. Patel. 1988. Near-wall turbulence models for complex flows including separation. AIAA J. 26(6):641-648. Chen, Q. and Z. Jiang. 1992. Significant questions in predicting room air motion. ASHRAE Transactions 98:929-939. Chen, Q., A. Moser and A. Huber. 1990. Prediction of buoyant, turbulent flow by a low-Reynolds-number k-8 model. ASHRAE Transactions 96(l):564-573. Chen, Q., A. Moser and P. Suter. 1992. A numerical study of indoor air quality and thermal comfort under six kinds of air diffusion. ASHRAE Transactions 98(1):203-217.

Chieng, C. C. and B. E. Launder. 1980. On the calculation of turbulent heat transport downstream from an abrupt pipe expansion. Numerical Heat Transfer 3:189-207. Choi, H. L., L. D. Albright, M. B. Timmons and Z. Warhaft. 1988. An application of the k-epsilon turbulence model to predict air distribution in a slot-ventilated enclosure. Transactions of the A^A^" 31(6): 1804-1814. Djilali, N., I. Gartshore and M. Salcudean. 1989. Calculation of convective heat transfer in recirculating turbulent flow using various near-wall turbulence models. Numerical Heat Transfer, Part A 16:189-212. Fang, J. B. and R. A. Grot. 1990. Numerical simulation of the performance of building ventilation systems. ASHRAE Transactions 96(2): 361-366. Gan, H. and H. Awbi. 1994. Numerical simulation of the indoor environment. Building and Environment 29(4):449-459. Hoff, S. J. 1990. Three-dimensional air speed and temperature distributions in a scaled-model livestock confinement facility. Unpub. Ph.D. diss., Univ. of Minnesota, Minneapolis. Knappmiller, K. D. and A. T. Kirkpatrick. 1994. Computational determination of the behavior of a cold air ceiling jet in a room with a plume. ASHRAE Trans. 100(2):677-684. Koestel, A., P. Hermann and G. Tuve. 1950. Comparative study of ventilating jets from various types of outlets. Transactions ASHVE56:A59-An%. Jones, P. J. and G. E. Whittle. 1992. Computational fluid dynamics for building air flow prediction - Current status and capabilities. Building and Environment 27:321-338. Lam, C. K. and K. Bremhorst. 1981. A modified form of the k-e model for predicting wall turbulence. Transactions ofASME 103:456-460. Launder, B. E. and D. B. Spalding. 1974. The numerical computation of turbulent flows. Computer Methods in Applied Mech. andEng. 3:269-289. Launder, B. E. and W. Rodi. 1981. The turbulent wall jet. Prog. Aerospace Sci. 19:81-128. . 1983. The turbulent wall jet-measurements and modeling. AAIW. Rev. Fluid Mech. 15:429-459. Liu, Q., S. J. Hoff and D. S. Bundy. 1995a. Advances in computational fluid dynamics: A review from a ventilation simulation perspective. ASAE Paper No. 95-4476. St. Joseph, Mich.: ASAE. . 1995b. A comparison of three proposed ventilation criteria. ASAE Paper No. 95-4642. St. Joseph, Mich.: ASAE. Maghirang, R. G. and H. B. Manbeck. 1993. Modeling particle transport in slot-inlet ventilated airspaces. Transactions of the A&4: 36(5): 1449-1459. Murakami, S., S. Kato and R. Ooka. 1994. Comparison of numerical predictions of horizontal nonisothermal jet in a room with three turbulence models - k- EVM, ASM, and DSM. ASHRAE Transactions 100(2): 697-704. Murakami, S., S. Kato and H. Nakagawa. 1991. Numerical prediction of horizontal nonisothermal 3-D jet in room based on the k-e model. ASHRAE Transactions 97:38-48. Myers, G. E., J. J. Schauer and R. H. Eustis. 1963. Plane turbulent wall jet flow development and friction factor. J. Basic Eng., Transactions ofASME 85:47-54. Nallasamy, M. 1987. Turbulence models and their applications to the prediction of intemal flows: A review. Computers & Fluids 15(2): 151-194. Narasimha, R., K. Y. Narayan and S. P. Parthasarathy. 1973. Parametric analysis of turbulent wall jets in still air. Aeronautical J. 77:354-359. Ng, K. H. and D. B. Spalding. 1972. Turbulence model for boundary layers near walls. The Physics of Fluids 15(1): 20-30. Patankar, S. V. 1980. Numerical Heat Transfer and Fluid Flow. New York: McGraw-Hill. Patel, V. C , J. T. Chon and J. Y. Yoon. 1991. Turbulent flow in a channel with a wavy wall. J. Fluids Eng. 113:579-586.

VOL. 39(2):689-698

697

Patel, V. C , W. Rodi and G. Scheuerer. 1984. Turbulence models for near-wall and low Reynolds number flows: A review. AMA 7. 23:1038-1319. Rajaratnam, N. 1976. Turbulent Jets. Amsterdam, The Netherlands: Elsevier. Rodi, W and D. B. Spalding. 1970. A two-parameter model of turbulence, and its application to free jets. Warme- und Stojfubertragung 3:85-95. Setrak, A. A. 1988. The effect of rectangular obstacles on the diffusion of wall jet. Ph.D. thesis, Napier Polytechnic, U.K. Schwarz, W. H. and W. P. Cosart. 1961. The two-dimensional turbulent wall-jet. 7. Fluid Mech. 10:481-495. Sigalla, A. 1958. Measurements of skin friction in a plane turbulent wall jet. J. of the Royal Aeronautical Soc. 62:873-877. Thangam, S. and C. G. Speziale. 1992. Turbulent flow past a backward-facing step: A critical evaluation of two-equation models. AMAy. 30:1314-1320. Tuve, G. L. 1953. Air velocities in ventilating jets. Transactions ASHVE 59:261-279.

Walker, J. N. 1977. Review of the theoretical relationships of isothermal ventilating air jets. Transactions oftheASAE 20:517-522. White, F. 1991. Viscous Fluid Flow. New York: McGraw-Hill. Wilcox, D. C. 1993. Comparison of two-equation turbulence models for boundary layers with pressure gradient. AIAA J. 31:1414-1421. Wilson, J. D., M. L. Esmay and S. Persson. 1970. Wall-jet velocity and temperature profiles resulting from a ventilation inlet. Transactions oftheASAE 13(2):77-81. Worley, M. S. and H. B. Manbeck. 1995. Modeling particle transport and air flow in ceiling inlet ventilation systems. Transactions oftheASAE 38(l):231-239. Wygnanski, I., Y. Katz and E. Horev. 1992. On the applicability of various scaling laws to the turbulent wall jet. J. of Fluid Mech. 234:669-690.

698

TRANSACTIONS OF THE ASAE

S-ar putea să vă placă și