You are on page 1of 4

The seventh International Conference on Urban Climate,

29 June - 3 July 2009, Yokohama, Japan

APPLICABILITY OF CFD PREDICTION TO THREE-DIMENSIONAL SNOWDRIFT


AROUND A CUBIC BUILDING MODEL
Yoshihide Tominaga*, Tsubasa Okaze**, Akashi Mochida**, Hiroshi Yoshino**, Yu Ito**
*Niigata Institute of Technology, Kashiwazaki, Japan; **Tohoku University, Sendai, Japan

Abstract

This paper presents the results of CFD prediction of snowdrift around a cubic building model, in which a new
snowdrift model developed by the present authors is examined. Numerical results using this new model are
compared with results obtained by the previous model and field measurements. It is confirmed that the predicted
snowdrift patterns, i.e., erosion around the upwind corners and deposition in front of and behind the model
obtained from the new CFD method, correspond well with those of the field measurements

Key words: CFD, Snowdrift, Turbulent flow

1. INTRODUCTION

Snowdrift formation around buildings is a major problem in snowy regions. Wind and water tunnels using modeled
snow particles have often been used to predict it. However, these apparatuses are not easily adaptable to
snowdrift and have limitations with regard to similarity law. Therefore, CFD (Computational Fluid Dynamics)
methods have been investigated as design tools for snowdrift around buildings in snowy regions.
This paper presents the results of CFD prediction of snowdrift around a cubic building model using a new
snowdrift model based on experimental and numerical studies by the authors. This model includes the effect of
buoyancy due to local difference in drifting snow densities and turbulence dissipation due to snow particles. The
numerical results are compared with data obtained from detailed field measurements taken to validate the
accuracy of the snowdrift modeling.

2. OUTLINES OF COMPUTATIONS

2.1. Flowfield Analyzed

Snowdrift around a surface mounted cubic model employed in detailed field measurements carried out by Oikawa
et al.(1999) in Sapporo, Hokkaido was adopted as the analyzed flowfield. The model was 1.0[m] high, which
corresponds to the actual size of the cube used in the field measurement.

2.2. Outline of snowdrift model

The governing equations of the new snowdrift model proposed in this study are shown in Table 1. The modified k-
ε model proposed by Durbin (1996) is used as the turbulence model. The underlined terms are added to take into
account the effects of snow on the flowfield. The underlined terms with ‘A’ and ‘B’ express the buoyancy effects of
local difference in drifting snow densities and the effects of snow particles on the turbulence property, respectively.
In the snowdrift modeling of most recent studies, including the previous model by the authors (Tominaga
et al., 2006), saltation and suspension of snow particles were modeled separately. The model proposed by
Pomeroy and Gray (1990) for the transport rate of drifting snow in a saltation layer has usually been used in these
studies. However, as demonstrated by the authors (Okaze et al., 2008), this model is applicable only to an
equilibrium saltation layer, not to a non-equilibrium region like flow around buildings. Therefore, a new approach
to evaluating the transport rate on drifting snow in a saltation layer was developed. In the approach adopted here,
the transport equation for drifting snow density <Φ> is solved without distinguishing between suspension and
saltation. This idea is based on the assumption that horizontal saltation velocity equals horizontal wind velocity.
∂ Φ
+
∂ Φ uj
+
∂ Φ wf
=

⎡− u ' Φ ' ⎤
(15)
∂x j ⎣ ⎦
j
∂t ∂x j ∂x3
The drift density flux is modeled to include the buoyancy effects of density difference as follows:
ν ∂ Φ k ρ − ρa (16)
u 'Φ ' = − t
j − C g s s3 Φ ′2 δ j3
σ s ∂x j ε ρ s ρa
2
k ν ⎛∂ Φ ⎞ (17)
Φ ′2 = − Cs 4 t ⎜ ⎟
ε σ s ⎝ ∂x3 ⎠
Here, CS3=0.25, CS4=1.6
Snowfall velocity <wf> [m/s] is defined as the value on the snow surface (except when wind velocity
equals zero). Therefore, <wf><Φp>[kg/m2s] corresponds to snow deposition flux on the surface. Here, <Φp> is the
drift density at the first grid adjacent to the snow surface in the CFD calculation selected as the CV. Subscript p
means a CV value. The deposition rate Mdep [kg/s] on the horizontal surface of the CV is given by:
The seventh International Conference on Urban Climate,
29 June - 3 July 2009, Yokohama, Japan

Table 1 Proposed governing equations for revised k-ε model (Okaze et al., 2008)
Continuity equation
∂ uj
=0 (1)
∂x j
Reynolds equation
∂ ui ∂ ui u j 1 ∂ p ⎛ ∂ ui ⎞
∂ ρ − ρa
+ =− ⋅ +⎜⎜ν ⎟⎟ −g s Φ δi 3 (2)
∂t ∂x j ρa ∂xi ∂x j
⎝ ∂x j ⎠ ρs ρa A
ρa : density of air, ρS : density of snow
Transport equation of k
∂k ∂k ∂ ⎛ ν t ∂k ⎞
+ uj = ⎜ ⎟ + Pk +Gk − ε + S k ) (3)
∂t ∂x j ∂x j ⎜⎝ σ k ∂x j ⎟⎠ A B

Transport equation of ε
∂ε
∂t
+ uj
∂ε
=
∂ ⎛ ν t ∂ε ⎞ ε

∂x j ∂x j ⎜⎝ σ ε ∂x j ⎟⎠ k
(
⎟ + Cε 1Pk − Cε 2ε +Cε 3Gk  
A
+ Sε
B ) (4)

ρ s − ρa
Pk = ν t kT (5) Gk = − g ui′Φ ′ δ i 3 (6)
ρ s ρa
T = min(TS , TD ) (7) Gk>0 (Unstable):C3=C1=1.44
Gk<0(Stable):C3=0.0
2
k 1 1 ⎛ ∂ ui ∂ uj ⎞
TS = (8), TS = (9), S= ⎜ + ⎟ (10)
ε Cμ S 3 2 ⎜ ∂x j ∂xi ⎟
⎝ ⎠
Extra terms for reproducing the effects of snow particles
k ε
Sk = −Cks fs Φ (11) Sε = −Cε s f s Φ (12)
ρst * ρst *
α
⎡ ⎧ ⎫ ⎤
⎢ ⎪ t* ⎪ ⎥ D 2ρ
fs = ⎢1 − exp ⎨− ⎬ ⎥ (13) t* = S s (14)


⎪ As k ε ⎪ ⎥
⎩ ⎭ ⎦ ( ) 18μ

CkS=2.0, CεS=0, AS=10.0, DS : Diameter of snow particle [m]

M dep = − Φ p w f ΔxΔy (18)


2
where, ΔxΔy [m ] is the horizontal area of the CV.
The erosion rate Mero [kg/s] due to the shear stress on the horizontal surface of the CV is given by
eq.(19) (Anderson and Haff, 1988),
⎛ ut* ⎞
2
(19)

M ero = − Aρi u * 1 −⎟ ΔxΔy
⎜ u *s
2

⎝ ⎠
-4
where, A=5.0x10 , ρi: density of ice
As described above, the net deposition rate Mtotal [kg/s] on the horizontal surface of the CV is given by:
M total = M dep + M ero (20)
Variation of snow depth per unit time Δzs[m/s] is obtained from Mtotal divided by accumulated snow
density ρs [kg/m3] and ΔxΔy.
M total (21)
Δzs =
ρ s ΔxΔy
In addition, when erosion occurs ( u* s > u* t ), the surface boundary condition for the transport equation
of <Φ>is given by:
v ⎛∂ Φ ⎞ M (22)
− t ⎜ ⎟ = ero
σ s ⎝ ∂x3 ⎠ surface ΔxΔy

2.3. Computational cases

The computational cases conducted here are summarized in Table 2. The drift density <Φ> at the inflow boundary
and the threshold friction velocity are changed as parameters. The profile of drift density <Φ> at inflow for Case-3
is illustrated in Fig. 1. It should be noted that the inflow density of drifting snow should be set as constant with
height as long as the influence of saltation is not considered in the transport equation of <Φ>, as was done in the
previous study (Tominaga et al., 2006). The parameters relating to the physical properties of snow are given by
the preliminary study for boundary layer flow by the authors (Okaze et al., 2008) as shown in Table 3.
The seventh International Conference on Urban Climate,
29 June - 3 July 2009, Yokohama, Japan

Accumulated snow density ρS is set to correspond to the value observed in the field measurement (Oikawa et al.,
1999).
Firstly, the steady flowfield was calculated without drifting snow, and then the unsteady computation was
conducted considering drifting snow.
Table 2 Computational cases
<Φ> at inflow boundary [kg/m3] Threshold friction velocity <u*t> [m/s]
Case-1 0.05 0.21
Case-2 0.05 0.15
cf. Fig.1
Case-3 0.15
(0.05 at upper boundary)

1.5 Case1, 2
Case 3
z/H

0.5

0
0 0.1 0.2 0.3 0.4 0.5 0.6
[kg/m3]
Fig. 1 Inflow profiles of the drift density (Case-3)
Table 3 Parameters relating to physical properties of snow
Accumulated snow
Snowfall velocity <wf> 0.2[m/s] 150 [kg/m3]
density ρS
Diameter of snow -4
1.5x10 [m] z0 at snow surface 3.0×10-5 [m]
particle DS

3. RESULTS AND DISCUSSION

3.1. Influence of boundary condition on snow depth

The horizontal distributions of snow depth at t*=100 for all cases using the new snowdrift model are shown in Fig.
2. Here, t* is non-dimensional time defined by H and <uH>. The deep-colored part in the figure indicates large
snow coverage. The snow depths are normalized by the value at the reference point. Only for Case-2, the ideal
snow depth, obtained by the boundary condition (t<wf><Φ>/ρs), is made dimensionless because no snow
accumulation occurs far from the building. In Case-1 (Fig. 2(1)), there are small undulations of snow over the
whole domain, although small erosion regions are formed near the upwind corners of the building. This is
because excess shear stress occurs only at the separation regions around the building due to the large threshold
friction velocity in this case (cf. Table 2). However, in Case-2, the snow is eroded over a wide region, except for
the area in front of and behind the building, due to the small <u*t> value. Where <u*s> at inflow is smaller than
<u*t>, the vertical profile of <Φ> at inflow can be set as constant with height because saltation does not occur.
However, where <u*s> at inflow is larger than <u*t>, such as for Case-2, the large value of <Φ> due to saltation
near the snow surface at inflow must be considered. For this reason, Case-2 underestimates the snow deposition
over the whole domain due to this underestimation of incoming drift density.
In Case 3, which has large drift density near snow surface at inflow due to saltation (cf. Fig. 1), the
depositions in front of the building and outside the separation regions become large. As a result, two large erosion
regions are clearly observed near the upwind corners of the building.

Wind
風向 Wind
風向 Wind
風向

0 2.0 0 2.0 0 2.0


(1) Case-1 (2) Case-2 (3) Case-3
Fig. 2 Horizontal distribution of normalized snow depth
The seventh International Conference on Urban Climate,
29 June - 3 July 2009, Yokohama, Japan

3.2. Comparison of snow depth distribution with field measurement results

Fig. 3 compares the snow depth distributions obtained from the previous CFD model (Tominaga et al., 2008) and
the present model with those obtained from the field measurement. Snow depth is normalized by the value at the
reference point, which was not affected by the building. The erosion regions near the upwind corners of the
building in the present CFD result are much larger than the previous results. Furthermore, the large peaks of
snow accumulation outside the separation regions near the building, which were observed in the previous model,
are not seen in the field measurement. This over-prediction of snow deposition is mainly caused by the fact the
Pomeroy’s model overestimates the transport rate in a non-equilibrium saltation layer such as flow around a
building. The result of the snowdrift patterns, i.e., erosion around the upwind corners and deposition in front of
and behind the building, obtained by the present CFD shows good correspondence with the field measurements.
Fig. 4 compares the horizontal distributions along a lateral line crossing the building. The results of the previous
CFD have too large peaks around the building, which are not observed in the present CFD results and the field
measurement.

2
Wind
風向
Wind Wind 1.8

1.6 0.0
0.4
1.4

1.2 0.8
1
1.2
0.8

0.6

0.4
0.4
0.2
1.2
0

(1) Previous method (2) Present method (Case-3) (3) field measurement
Fig. 3 Comparison of horizontal distribution of normalized snow depth

5 Field measurement
Normalized snow depth

Wind Present CFD


4 Previous CFD(9)
3
2
1
0
-2 -1.5 -1 -0.5 0 0.5 1 1.5 y/H 2
Fig. 4 Comparison between normalized snow depths obtained from CFD and field measurements along lateral
line (x/H=0.1)

4. CONCLUSIONS

The accuracy of the CFD method for predicting snowdrift around a cubic building model using the new snowdrift
model proposed by the authors was was examined by comparing its results with those from the previous CFD
model and field measurements. The snowdrift patterns, i.e., erosion around the upwind corners and deposition in
front of and behind the building obtained by the present CFD shows better correspondence with those obtained
from the field measurement and the wind tunnel experiment than the previous CFD results.

References
Anderson, R. S. and Haff, P. K. (1988), “Simulation of eolian saltation”, Science, 241, 820-823.
Durbin, P.A. (1996), “On the k-ε stagnation point anomaly”, Int. J. Heat and Fluid Flow, 17, 89-90.
Okaze, T., Mochida, A., Tominaga, Y., Nemoto, M., Ito, Y., Shida, T., Sato, T., Yoshino, H. (2008), “Modeling of
Drifting Snow Development in a boundary leyer and its effect on wind field”, 6th Snow Engineering Conference,
Whistler, B.C., Canada, June 1-5.
Oikawa, S., Tomabechi, T., Ishihara, T. (1999), “One-day Observations of Snowdrifts Around a Model Cube”, J. of
Snow Eng. of Japan, 15(4), 3-11 (in Japanese).
Pomeroy, J. W., and Gray, D, M. (1990), “Saltation of snow”, Water Resource Research, 26(7), 1583-1594.
Tominaga, Y., Mochida, A., Yoshino, H., Shida, T., Okaze, T. (2006), “CFD Prediction of Snowdrift around a
Cubic Building Model”, The Fourth International Symposium on Computational Wind Engineering (CWE2006),
Yokohama, Japan.

You might also like