Pinard 2001

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

1762 JOURNAL OF APPLIED METEOROLOGY VOLUME 40

First- and Second-Order Closure Models for Wind in a Plant Canopy


J. D. JEAN-PAUL PINARD AND JOHN D. WILSON
Department of Earth and Atmospheric Sciences, University of Alberta, Edmonton, Alberta, Canada

(Manuscript received 19 October 2000, in final form 26 February 2001)

ABSTRACT
Katul and Chang recently compared the performance of two second-order closure models with observations
of wind and turbulence in the Duke Forest canopy, noting that such models ‘‘alleviate some of the theoretical
objections to first-order closure.’’ This paper demonstrates that, notwithstanding those (valid) theoretical objec-
tions, Duke Forest wind simulations of comparable quality can be obtained using a first-order closure, namely,
eddy viscosity K } lÏk, where k is the turbulent kinetic energy and l is a turbulence length scale. It is concluded
that, most often, uncertainty in the drag coefficient will limit the accuracy of modeled wind statistics, regardless
of the turbulence closure chosen.

1. Introduction is the leaf area density], and ] p /] x is (any) streamwise


gradient in kinematic pressure. Wind models are distin-
As a precursor to rational scientific treatment of many guished as being of first- or second-order closure, ac-
forest processes, such as aerial spray dispersion or the cording to their means of providing the Reynolds stress
spread of pollen or plant pathogens, profiles of mean in Eq. (1).
horizontal wind velocity u and turbulence statistics (e.g., Seeking ‘‘a practical framework for computing need-
component wind velocity variances s 2u , s 2y , and s 2w) need ed velocity statistics for modeling scalar transport,’’ Ka-
to be known. Furthermore these statistical properties of tul and Chang (1999, hereinafter KC99) compared two
the wind modulate the relationship between the mean second-order closure models (Wilson and Shaw 1977,
properties of the airstream (e.g., humidity and carbon hereinafter WS77; and Wilson 1988, hereinafter W88)
dioxide concentration) and the rates of emission/ab- with measured winds in the Duke Forest of North Car-
sorption by the vegetation (leaf transpiration and pho-
olina. They concluded that WS77 produced a slightly
tosynthesis). Thus atmosphere–forest interactions are of
better mean velocity profile u but a worse standard de-
great importance in such fields as agronomy and for-
viation profile s u than did W88 and that the modeled
estry.
third moments were inconsistent with measurements.
Dynamical models for wind in a ‘‘horizontally uni-
Our point of departure with respect to KC99 is that,
form’’ plant canopy, in which by definition flow statis-
whereas these authors sought to establish which was the
tics may vary with the height z but not along the hor-
better second-order model of the two they compared,
izontal coordinates, are based on a mean streamwise
they did not enquire whether a first-order closure might
momentum equation,1
have been comparably suitable. In raising this question,
]u9w9 ]p we do not dispute the logical superiority of second-order
5 2Cd au 2 2 , (1) closures or overlook their promise to partition the tur-
]z ]x
bulent kinetic energy k 5 1/2(s 2u 1 s 2y 1 s 2w) into its
where u9w9 is the mean vertical flux of streamwise mo- components (s 2w, etc.); we merely wish to highlight am-
mentum (Reynolds stress), C d au 2 parameterizes drag on biguities or uncertainties common to both approaches
plant parts [C d 5 C d (z) is the drag coefficient, a 5 a(z) (e.g., optimal closure constants, means of providing the
drag coefficient) and query whether these imply that the
vaunted superiority of second-order closure is incon-
1
For a derivation see Wilson and Shaw (1977) or Raupach and
Shaw (1982). In simplifying to obtain Eq. (1) we assumed the dis- sequential. Thus our goal is to establish whether there
persive momentum flux to be negligible. is any striking gain in accuracy of modeling the first
and second velocity moments when the more complex
approach (second-order closure) is taken.
Corresponding author address: J. D. Jean-Paul Pinard, Department
of Earth and Atmospheric Sciences, 1-26 Earth Sciences Building, In what follows, we briefly review an existing first-
University of Alberta, Edmonton, AB T6G 2E3, Canada. order closure model for canopy flows (that of Wilson
E-mail: jpinard@ualberta.ca et al. 1998, hereinafter WFR98) and our numerical

q 2001 American Meteorological Society


OCTOBER 2001 PINARD AND WILSON 1763

method for solving it. In section 3, we discuss the de-


termination of the drag coefficient from measurements.
In section 4, we compare our own simulations of the
Duke Forest winds with the observations and simula-
tions of KC99.

2. A first-order closure model for wind in a


uniform canopy
A balance between the divergence of the mean ver-
tical flux of streamwise momentum and drag on plant
parts controls the airflow within a uniform canopy [Eq.
(1)]. In a first-order (or ‘‘flux-gradient’’) closure, tur-
bulence is approximated as being equivalent to an in-
creased viscosity of the fluid; that is, the shear stress is
modeled as u9w9 5 2 K(]u /]z 1 ]w /]x), where the
eddy viscosity K is determined by the flow itself. Our
present assumption of a uniform canopy eliminates the
term in mean vertical velocity.
Higher-order closures, such as those exploited by
KC99, use the Navier–Stokes equations to obtain exact
governing equations for u9w9 and other turbulence sta-
tistics—equations that are subsequently simplified so as
to attain closure (as many equations as unknowns). FIG. 1. The plot for (finite-difference based) C d (z) is highly variable
and shows how sensitive it can be to noise in the shear stress gradient.
Many mechanisms are thereby resolved exactly, for ex- The calculated (polynomial based) drag coefficient C dp (z) is negative
ample, the shear production u9w9 (]u /]z) and advection at the ground and positive above z 5 2 m. The value for the height-
u (]u92 /]x) of u variance; other mechanisms have to be averaged drag coefficient Cd , which is the same for both our cal-
approximated (for example, the triple moments u9u9u9 i j k
culation and that of KC99, is shown for comparison.
are often assumed to be fluxes driven by spatial gra-
dients in u9u9
i j ). Substitution of a higher-order closure
for a first-order closure is a step in the direction of local equilibrium layer far above the canopy (and is the
greater rigor. However, both types of models require standard parameterization for viscous dissipation). The
empirical inputs that are somewhat uncertain (e.g., the form drag «fd represents conversion of resolved TKE to
drag coefficient), and it is our contention and the point small, rapidly dissipated ‘‘wake scales,’’ and the closure
of this paper that for many purposes the first-order clo- constant a was reasoned by WFR98 to be approximately
sure model will be sufficient. equal to 1.
Crucial to the WFR98 closure is the specification of
a. Equations determining the eddy viscosity the length scale l, which is characterized as the max-
imum of inner and outer length scales
Our calculations will use the first-order closure of
1 1 1
WFR98, who parameterized the eddy viscosity as K 5 5 1 , and (3a)
l (z)Ïc e k(z), where l (z) is an algebraic length scale li ky z lc
(defined below); c e is a (fairly well known) quasi con- 1 1 1
stant, giving the equilibrium shear-stress: turbulent ki- 5 1 . (3b)
netic energy (TKE) ratio (u*2 0 /k 0 ) immediately above the lo k y (z 2 d) L`
canopy; and k is the TKE, calculated from a simplified Here k y 5 0.4 is the von Kármán constant and d is the
transport equation displacement height (usually set as 2/3 of the canopy
height hc). The outer scale l 0 recognizes the flow dis-

1 2
]k ]u ] ]k placement by the canopy and, for the case of a wind-
5 0 5 2u9w9 1 mK 2 «. (2)
]t ]z ]z ]z tunnel experiment, a finite limiting value L` . The inner
scale l i recognizes the limitation on eddy size due to
Here we neglect buoyant production, because our cal-
proximity to ground and the presence of the canopy,
culations pertain only to the case of neutral stratification.
which is felt through an upper limit l c to the inner length
The effective diffusion coefficient for TKE, mK, is pro-
scale:
portional to the eddy viscosity, and, as in WFR98, we
set m 5 0.2. The viscous dissipation is written as « 5 21

1 2
]u
max(«cc , «fd ), where «cc 5 (c e k) 3/2 /l and « fd 5 aC d au k. l c 5 cÏk(h c ) . (4)
The term «cc balances shear production of TKE in the ]z hc
1764 JOURNAL OF APPLIED METEOROLOGY VOLUME 40

Equation (4) is based on the suggestion of Raupach et TABLE 1. Parameters of the Duke canopy experiments.
al. (1996) that the strength of the wind shear at the Property Symbol Value
canopy height is critical to eddy transport within a can-
Canopy height hc 14.0 m
opy; WFR98 determined that the closure constant c is Model height Lid 10hc
approximately equal to 1.0. Numerous other formula- Bulk drag parameter calcu- Cdb 5 Cd ahc 0.34
tions of the eddy viscosity within and above a plant lated in section 3c
canopy have been suggested. The current one has been Height-averaged drag coeffi- Cd 0.2
cient from section 3b and
tested over a wider variety of uniform canopy flows than as KC99
most and has been shown to perform very well for dis- Displacement height d/hc 0.67
turbed winds in forest clearings (Wilson and Flesch Equilibrium variances as giv- su,y ,w/u*0 2.08, 1.86, 1.22
1999). en by KC99
Closure coefficients c, a, m 1.0, 1.0, 0.2
Limit to length scale L` 1.5
b. Partitioning TKE into its components Grid points/iterations 100/100

An advantage in principle of a second-order model


is its ability to partition the TKE into its components
(s 2u, s 2y , s 2w) and thus to make specific predictions of
1 ]x2 ,
]p
flow statistics vital to canopy transport—especially 2Aun u J11
m
1 Auc u Jm 2 Aus u J21
m
5 Dz 2 (7)
s w (z), which is a necessary input to Lagrangian sto- J

chastic dispersion models such as have recently been where A un 5 K n /Dz, A su 5 K s /Dz, A uc 5 A un 1 A us 1
used to infer canopy source/sink strengths (e.g., for car- DzC d ah c | u m21
J | , and the superscript m designates the
bon dioxide) from measured concentration profiles. The mth iterative guess for the u field. Similar results obtain
K}lÏk closure does not offer this information, and the for the TKE equation, and solving these equations im-
best one can easily do is to assume constant partitioning plies the inversion of tridiagonal matrices.
ratios s 2u /k 5 g u2, s 2y /k 5 g y2, and s 2w /k 5 g w2 . The Grids for mean velocity and TKE were staggered,
partitioning constants satisfy g u2 1 g y2 1 g w2 5 2 and with the lowest u grid point falling on ground and the
allow us to write s u /u* 5 g u /Ïc e, and so on, where uppermost TKE grid point lying at z 5 10h c . Upper
u* is friction velocity. Note that the analytical second- boundary conditions were a prescribed value (21) for
order closure of Massman and Weil (1999) likewise as- the (normalized) momentum influx u9w9 /u*2 and (except
sumed ‘‘that s 2u, s 2y , s 2w are each proportional to (k) and in the case of a wind-tunnel simulation, for which we
that the proportionality constants are the same as those set ]k/]z 5 0) a prescribed (equilibrium) value for the
at the canopy top.’’ normalized TKE, namely, k/u*2 5 1/c e 5 4.64. At the
bottom boundary, z 5 0, the mean wind speed u 5 0,
and we imposed ]k/]z 5 0 (flux of TKE to ground
c. Numerical method used to implement the K-theory vanishes).
model
Discretization equations are formulated following the
d. Revisiting and confirming the Wilson et al. (1998)
approach of Patankar (1980). The governing equations
simulations
are integrated across a control layer spanning height
range s # z # n to yield a ‘‘neighbor equation,’’ linking We carried out new simulations of the same three
velocity at the Jth grid point to its values above (J 1 uniform canopy flows studied by WFR98 in developing
1) and below (J 2 1). Our momentum equation, closed the closure (details of these experiments can be obtained
using K theory, upon such integration gives from WFR98 and references therein). Briefly, ‘‘Furry
Hill’’ was a canopy of flexible strands of fishing line

[ ]
n
]u (h c 5 4.7 cm), stretching upwind and over a model hill
2K 5 2DzCJ u J2 , (5)
]z s
in a wind tunnel; measurements cited here are from the
region upwind from the hill. The ‘‘Tombstone Canopy’’
where C J is the product of the drag coefficient and area was a regular array of vertical bars in the same wind
density (see section 3). The momentum fluxes across tunnel, each tombstone being h c 5 6 cm high by 1 cm
the north (n) and south (s) faces are evaluated as in cross-stream width, with 6-cm cross-stream and 4.4-
cm alongstream spacing. The Elora field experiment
1K ]z 2 5 K
]u u J11 2 u J
, took place in a mature, uniform corn canopy of height
Dz
n
n h c 5 2.2 m at Elora, Ontario, Canada.
We found that grid independence and convergence

1K ]z 2 5 K
]u u J 2 u J21 were assured in these one-dimensional canopy flow sim-
, (6) ulations, when 100 (or more) iterations were performed
Dz
s
s with grid spacing Dz 5 h c /100. Our new simulations
and the resulting neighbor equation for velocity is compared very closely to those reported by WFR98; the
OCTOBER 2001 PINARD AND WILSON 1765

differences, which were minor, we attribute to our hav- Because of the irregularity of our calculated profiles
ing used higher resolution so as to attain grid indepen- of drag coefficients C d (z) and C dp (z), we discarded Eq.
dence. (9) in favor of the following methods.

3. Provision of the drag coefficient b. Height-averaged drag coefficient


The drag coefficient parameterizes the drag on the If we have untrustworthy (or nonexistent) information
surfaces of leaves and limbs and exerts a controlling on u9w9 but confidence in the profiles a(z) and u (z), then
influence on the wind and turbulence among the plants. we can introduce a height-averaged drag coefficient
Depending on the quantity and quality of information C d , defined by
given (the profiles of leaf area and of the flow variables
u and u9w9 ), the drag coefficient may be determined at
three levels of detail.
E0
hc
]u9w9
]z
dz 5 2t h c 1 t (0) 5 2t hc 1 0

a. Height-variable drag coefficient


In principle, the best way to calculate the drag co-
5 2Cd E 0
hc

a(z)u(z) 2 dz, (10)

efficient, from good-quality micrometeorological mea- where the kinematic stress t hc 5 2 (u9w9 )hc . Rearrange-
surements, would be to calculate C d (z) by rearranging ment gives
Eq. (1):
t hc t hc /u*2

E
Cd 5 5
E
]u9w9 ]p
2 2
[ ]
hc hc 2

]z ]x u(z)
a(z)u(z) 2 dz a(z) dz
Cd (z) 5 2
. (8) 0 u*
a(z)u (z) 0

1
5
E
If we neglect the pressure gradient (which is accept- . (11)

[ ]
hc 2
able for atmospheric flows) and normalize, we can trans- u(z)
a(z) dz
form the above to 0
u*

1 2 1 u* 2
] u9w9 u9w9
2 2D This approach avoids the differentiation of noisy
]z/h c u*2 2
u9w9 data. The value we calculate for the Duke Forest
Cd (z) 5 5 , (9) (C d 5 0.2; see Fig. 1) is identical to the drag coefficient
[ ] [ ]
2 2

12
u(z) z u(z)
h c a(z) D h a(z) cited by KC99, which may therefore be of the same
u* hc c u* provenance.
where the D operator implies a ‘‘finite difference’’ along
the height axis. c. Bulk drag coefficient
In the Duke Forest of KC99, mean wind and shear
stress were measured at six points in and above the If, furthermore, we have poor or nonexistent infor-
canopy. We calculated C d according to Eq. (9) from the mation on the leaf area density profile a(z), then we
measured data u /u*, u9w9 /u*2 , and a(z) at the five sensor have no choice but to use a bulk coefficient C d a , which
is defined by

E E
levels falling within the canopy. An irregular profile of
C d resulted (Fig. 1), because the observations of u9w9 / hc
]u9w9 hc

u*2 carry considerable uncertainty and are (here, as of- dz 5 2t h c 1 0 5 2Cd a u(z) 2 dz. (12)
0
]z 0
ten) given at an irregularly spaced set of points.
To alleviate this difficulty in the application of Eq. By rearranging, we may define a dimensionless bulk
(9), one might consider ‘‘smoothly fitting’’ continuous coefficient C db 5 h cC d a ; that is,
polynomial curves to the measured profiles of mean 1
Cdb 5
E1
wind and shear stress. Substitution of the resulting func- . (13)

2 12
1 2
tions into Eq. (9) will then define a polynomial ap- u z
proximation C dp (z) for the drag coefficient. In the cur- d
0
u* hc
rent case, when this was tried, the outcome (Fig. 1) was
unsatisfactory, giving rise to a negative drag coefficient Once known, this bulk drag coefficient may be utilized
C dp deep in the canopy. Evidently this procedure re- in a normalized momentum equation, namely,
]u9w9/u*2 h c ](p/u*2 h c )
1 2
quires arbitrary steps, such as the choice of order of the u
2

polynomial, in fitting what may be highly irregular pro- 5 2Cdb 2 . (14)


](z/h c ) u* ]x
files (e.g., the leaf area density profile reported for the
Duke Forest).
1766 JOURNAL OF APPLIED METEOROLOGY VOLUME 40

FIG. 2. Normalized mean wind speed u of the first-order model vs


FIG. 3. The shear stress u9w9 , from the first-order model vs observed
observed (dots). The thick solid line is first-order closure with height-
(dots). The thick solid line is first-order closure with height-averaged
averaged Cd 5 0.2; the thin solid line is with a bulk C db 5 0.34. The
Cd 5 0.2; the thin solid line is with a bulk C db 5 0.34. The long-
long-dashed line is W88; the short-dashed line is WS77 as simulated
dashed line is W88; the short-dashed line is WS77 as simulated by
by KC99.
KC99.

For the Duke Forest, Eq. (13) yields C db 5 0.34. We


do not plot C db alongside our other estimates of the drag (these being the values at z/h c 5 1.097 cited in Table
coefficient on Fig. (1), because C db has a different mean- 3 of KC99), implying c e 5 1/4.64.
ing and is imposed in a different momentum equation Figures 2–5 show results of our simulations of the
than are C d (z), C dp (z), and C d . Duke Forest; the two second-order simulations of KC99,
namely WS77 and W88, have been overlaid onto the
current (first-order) results for easy comparison.
4. Comparing the first-order closure model with
Our normalized mean wind speed profile, Fig. 2,
Duke Forest observations and the second-order
closely matches the observed profile of KC99. This re-
simulations of Katul and Chang (1999)
sult is not very surprising, given that we derived the
Because our objective was to assess the WFR98 first- drag coefficients (both height averaged and bulk) by
order model against the second-order models examined integration of the observed mean wind speed profile. It
by KC99, in modeling the Duke Forest we followed should be expected, then, that the models should repro-
KC99 as closely as possible (see Table 1): the top of duce the mean wind and shear stress observations well.
the model domain was chosen as 10h c ; the height-av- If, on the basis of Fig. 2 we are to say that the models
eraged drag coefficient was C d 5 0.2 (same as KC99), are ‘‘good,’’ then perhaps the second-order models are
or, where we used a bulk drag coefficient, its value was not as good as the first-order model. The ‘‘noisiness’’
C db 5 0.34; the zero-plane displacement d/h c 5 0.67; of the second-order solutions probably results from the
and the normalized standard deviations s u /u* , s y /u* , absence of explicit diffusion in the momentum Eq. (1).
and s w /u* were set respectively as 2.08, 1.86, and 1.22 The first-order model also closely matches the mea-
OCTOBER 2001 PINARD AND WILSON 1767

FIG. 4. The turbulent kinetic energy k of the first-order model vs FIG. 5. Observed (dots) and modeled profiles of the normalized
observed. The solid-dot symbols are the TKE recalculated from s u , standard deviation of vertical velocity s w /u* . The thick solid line is
s y , and s w given by KC99. The hollow-dot symbols are TKE given first-order closure with height-averaged drag coefficient Cd 5 0.2;
by KC99. The thick solid line is first-order closure with height-av- the thin solid line is with a bulk drag coefficient C db 5 0.34. The
eraged C d 5 0.2; the thin solid line is with a bulk Cd 5 0.34. The long-dashed line is W88; the short-dashed line is WS77 as simulated
long-dashed line is W88; the short-dashed line is WS77 as simulated by KC99.
by KC99. Note that our specification c e 5 (k/u*2) 21
h c 5 1/4.64 requires
our modeled k profile to run through the observation at z/h c ø 1.
just as does our Eq. (3), the TKE equation of WS77
sured stress profile of KC99 (see Fig. 3) and is similar reduces, above the canopy and provided the transport
to that obtained by KC99 using the W88 model. Of term is dropped, to k 3/2 } u*2 l 3]u /]z 5 constant. This
course, one cannot simultaneously do a good job of the is because ]u /]z } u* /(z 2 d) and l 3 } (z 2 d). Thus,
mean wind and a bad job of the shear stress, for it is the differing above-canopy TKE profiles reflect the fact
implied by the way in which the drag coefficient is that treatment of TKE transport has differed, and it may
derived that the modeled mean wind and shear stress be that our specification m 5 0.2 (recommended by
must be consistent. WFR98, though recognizing that m 5 1 is the more
Figure 4 indicates that the first-order model simulates frequent suggestion) has resulted in underestimation of
the profile of TKE reasonably satisfactorily but under- TKE transport.
estimates the magnitude of the above-canopy TKE gra- Figure 5 compares the observed vertical profile of s w
dient suggested by the observations. [Note that there is in Duke Forest with simulations. Because the first-order
a discrepancy between values of the (measured) velocity model underestimates TKE deep in the canopy, it also
variances for Duke Forest tabulated by KC99 and the underestimates the velocity variances, and thus s w .
TKE they have plotted on their Fig. 1. We plot both the
original and the corrected TKE on our Fig. 4.]
5. Conclusions
Except for the influence of the TKE transport term,
first- and second-order models both would generate a The preceding results (and those of WFR98) suggest
height-invariant TKE above the canopy; for example, that the simpler first-order closure model will often sim-
1768 JOURNAL OF APPLIED METEOROLOGY VOLUME 40

ulate the fundamental wind properties (mean wind Massman, W. J., and J. C. Weil, 1999: An analytical one-dimensional
speed, mean shear stress, and turbulent kinetic energy) second-order closure model of turbulence statistics and the La-
grangian time scale within and above plant canopies of arbitrary
of a canopy flow as well as a second-order closure model structure. Bound.-Layer Meteor., 91, 81–107.
will. The implication is that the theoretical superiority Patankar, S. V., 1980: Numerical Heat Transfer and Fluid Flow.
of a second-order model is moot, in the face of the large Hemisphere, 197 pp.
uncertainty with respect to the canopy drag coefficient Raupach, M. R., and R. H. Shaw, 1982: Averaging procedures for
that will ordinarily prevail, in any routine application flow within vegetation canopies. Bound.-Layer Meteor., 22, 79–
of these kinds of models. Thus, bearing in mind that for 90.
——, J. J. Finnigan, and Y. Brunet, 1996: Coherent eddies and tur-
two- and three-dimensional flows a second-order model bulence in vegetation canopies: The mixing-layer analogy.
is laborious, one ought not to overlook the competence Bound.-Layer Meteor., 78, 351–382.
of the simpler first-order model. Wilson, J. D., 1988: A second-order closure model for flow through
vegetation. Bound.-Layer Meteor., 42, 371–392.
Acknowledgments. Support from the Natural Sciences ——, and T. K. Flesch, 1999: Wind and remnant tree sway in forest
and Engineering Research Council of Canada is ac- cutblocks. Part III: A windflow model to diagnose spatial var-
knowledged. iation. Agric. For. Meteor., 93, 259–282.
——, J. J. Finnigan, and M. R. Raupach, 1998: A first-order closure
for disturbed plant canopy flows, and its application to windflow
REFERENCES through a canopy on a ridge. Quart. J. Roy. Meteor. Soc., 124,
Katul, G. G., and W.-H. Chang, 1999: Principal length scales in sec- 705–732.
ond-order closure models for canopy turbulence. J. Appl. Me- Wilson, N. R., and R. H. Shaw, 1977: A higher order closure model
teor., 38, 1631–1643. for canopy flow. J. Appl. Meteor., 16, 1197–1205.

You might also like