Skip to main content

Advertisement

Log in

Analytically Modelling Mean Wind and Stress Profiles in Canopies

  • Article
  • Published:
Boundary-Layer Meteorology Aims and scope Submit manuscript

An Erratum to this article was published on 31 May 2014

Abstract

An analytical model for mean wind profiles in sparse canopies (W. Wang, Boundary-Layer Meteorol 142:383–399, 2012) has been further developed, with (1) an explicit solution being derived, and (2) a linear term being added to the \(K\)-closure scheme to improve the shear-stress parametrization when the contribution of non-local transport is significant. Results from large-eddy simulations and from laboratory experiments are used to evaluate the model and adjust model parameters, showing that the model can well simulate canopy wind and stress profiles not only for sparse-canopy scenarios, but also for dense-canopy scenarios. The analytical solution converges exactly to the standard surface-layer logarithmic wind profile in the case of zero canopy density, and tends to an exponential wind profile for a dense canopy.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. A canopy is called a dense canopy when its frontal area index is \(>\)2.5 where wind speed is very small in the lower part of the canopy, according to Inoue (1963)’s exponential model. Otherwise, it is called a sparse canopy.

  2. Frontal area index can be assumed to be half of the commonly used single-sided leaf area index (LAI) for field plant canopies (e.g., Raupach et al. 1996).

  3. Under horizontally-homogenous conditions, the vertical gradient of stress is usually assumed to be balanced by canopy drag in all existing analytical canopy wind models (e.g. Inoue 1963; Cowan 1968), and therefore \(F_\mathrm{p}\) is not involved. Following this convention, \(F_\mathrm{p}\) is set to zero for a fair comparison with existing analytical canopy wind models.

References

  • Albini FA (1981) A phenomenological model for wind speed and shear stress profiles in vegetation cover layers. J Appl Meteorol 20(11):1325–1335

    Article  Google Scholar 

  • Ayotte KW, Finnigan JJ, Raupach MR (1999) A second-order closure for neutrally stratified vegetative canopy flows. Boundary-Layer Meteorol 90(2):189–216

    Google Scholar 

  • Bailey BN, Stoll R (2013) Turbulence in sparse, organized vegetative canopies: a large-eddy simulation study. Boundary-Layer Meteorol 147(3):369–400

    Article  Google Scholar 

  • Belcher SE (2005) Mixing and transport in urban areas. Philos Trans Math Phys Eng Sci 363(1837):2947–2968

    Article  Google Scholar 

  • Belcher SE, Jerram N, Hunt JCR (2003) Adjustment of a turbulent boundary layer to a canopy of roughness elements. J Fluid Mech 488:369–398

    Article  Google Scholar 

  • Belcher SE, Finnigan JJ, Harman IN (2008) Flows through forest canopies in complex terrain. Eco Appl 18(6):1436–1453

    Article  Google Scholar 

  • Brunet Y, Finnigan JJ, Raupach MR (1994) A wind tunnel study of air flow in waving wheat: single point velocity statistics. Boundary-Layer Meteorol 70:95–132

    Article  Google Scholar 

  • Bryan GH, Fritsch JM (2002) A benchmark simulation for moist nonhydrostatic numerical models. Mon Weather Rev 130(12):2917–2928

    Article  Google Scholar 

  • Cescatti A, Marcolla B (2004) Drag coefficient and turbulence intensity in conifer canopies. Agric For Meteorol 121(3–4):197–206

    Article  Google Scholar 

  • Cheng H, Castro IP (2002) Near wall flow over urban-like roughness. Boundary-Layer Meteorol 104(2):229–259

    Article  Google Scholar 

  • Cionco RM (1965) A mathematical model for air flow in a vegetative canopy. J Appl Meteorol 4(4):517–522

    Article  Google Scholar 

  • Coceal O, Belcher SE (2004) A canopy model of mean winds through urban areas. Q J R Meteorol Soc 130(599):1349–1372

    Article  Google Scholar 

  • Cowan IR (1968) Mass, heat and momentum exchange between stands of plants and their atmospheric environment. Q J R Meteorol Soc 94:523–544

    Article  Google Scholar 

  • Deardorff JW (1966) The counter-gradient heat flux in the lower atmosphere and in the laboratory. J Atmos Sci 23:503–506

    Article  Google Scholar 

  • Deardorff JW (1980) Stratocumulus-capped mixed layers derived from a 3-dimensional model. Boundary-Layer Meteorol 18(4):495–527

    Article  Google Scholar 

  • Denmead OT, Bradley EF (1985) Flux–gradient relationships in a forest canopy. In: Hutchison BA, Hicks BB (eds) The forest-atmosphere interaction. D. Reidel, Dordrecht, pp 421–442

    Chapter  Google Scholar 

  • Di Sabatino S, Solazzo E, Paradisi P, Britter R (2008) A simple model for spatially-averaged wind profiles within and above an urban canopy. Boundary-Layer Meteorol 127(1):131–151

    Article  Google Scholar 

  • Dwyer MJ, Patton EG, Shaw RH (1997) Turbulent kinetic energy budgets from a large-eddy simulation of airflow above and within a forest canopy. Boundary-Layer Meteorol 84(1):23–43

    Article  Google Scholar 

  • Finnigan JJ (1985) Turbulent transport in flexible plant canopies. In: Hutchison BA, Hicks BB (eds) The forest–atmosphere interaction. D. Reidel, Dordrecht, pp 443–480

    Chapter  Google Scholar 

  • Finnigan JJ (2000) Turbulence in plant canopies. Annu Rev Fluid Mech 32(1):519–571

    Article  Google Scholar 

  • Finnigan JJ, Belcher SE (2004) Flow over a hill covered with a plant canopy. Q J R Meteorol Soc 130(596):1–29

    Article  Google Scholar 

  • Finnigan JJ, Shaw RH (2000) A wind tunnel study of airflow in waving wheat: an EOF analysis of the structure of the large-eddy motion. Boundary-Layer Meteorol 96:211–255

    Article  Google Scholar 

  • Finnigan JJ, Shaw RH, Patton EG (2009) Turbulence structure above a vegetation canopy. J Fluid Mech 637:387–424

    Article  Google Scholar 

  • Frech M, Mahrt L (1995) A two-scale mixing formulation for the atmospheric boundary layer. Boundary-Layer Meteorol 73(1):91–104

    Article  Google Scholar 

  • Harman I, Finnigan J (2007) A simple unified theory for flow in the canopy and roughness sublayer. Boundary-Layer Meteorol 123(2):339–363

    Article  Google Scholar 

  • Holtslag AAM, Moeng C-H (1991) Eddy diffusivity and countergradient transport in the convective atmospheric boundary layer. J Atmos Sci 48:1690–1698

    Article  Google Scholar 

  • Inoue E (1963) On the turbulent structure of air flow within crop canopies. J Meteorol Soc Jpn 41:317–326

    Google Scholar 

  • Kang S-L, Davis KJ (2008) The effects of mesoscale surface heterogeneity on the fair-weather convective atmospheric boundary layer. J Atmos Sci 65(10):3197–3213

    Article  Google Scholar 

  • Kono T, Tamura T, Ashie Y (2010) Numerical investigations of mean winds within canopies of regularly arrayed cubical buildings under neutral stability conditions. Boundary-Layer Meteorol 134(1):131–155

    Article  Google Scholar 

  • Landsberg JJ, James GB (1971) Wind profiles in plant canopies: studies on an analytical model. J Appl Ecol 8(3):729–741

    Article  Google Scholar 

  • Macdonald RW (2000) Modelling the mean velocity profile in the urban canopy layer. Boundary-Layer Meteorol 97(1):25–45

    Article  Google Scholar 

  • Massman WJ (1997) An analytical one-dimensional model of momentum transfer by vegetation of arbitrary structure. Boundary-Layer Meteorol 83(3):407–421

    Article  Google Scholar 

  • Nepf HM (2012) Flow and transport in regions with aquatic vegetation. Annu Rev Fluid Mech 44(1):123–142

    Article  Google Scholar 

  • Novak MD, Warland JS, Orchansky AL, Ketler R, Green S (2000) Wind tunnel and field measurements of turbulent flow in forests. Part I: uniformly thinned stands. Boundary-Layer Meteorol 95(3):457–495

    Article  Google Scholar 

  • Patton EG, Katul GG (2009) Turbulent pressure and velocity perturbations induced by gentle hills covered with sparse and dense canopies. Boundary-Layer Meteorol 133(2):189–217

    Article  Google Scholar 

  • Pietri L, Petroff A, Amielh M, Anselmet F (2009) Turbulence characteristics within sparse and dense canopies. Environ Fluid Mech 9(3):297–320

    Article  Google Scholar 

  • Pinard JDJ-P, Wilson JD (2001) First- and second-order closure models for wind in a plant canopy. J Appl Meteorol 40(10):1762–1768

    Article  Google Scholar 

  • Poggi D, Katul GG, Albertson JD (2004a) Momentum transfer and turbulent kinetic energy budgets within a dense model canopy. Boundary-Layer Meteorol 111(3):589–614

    Article  Google Scholar 

  • Poggi D, Porporato A, Ridolfi L, Albertson JD, Katul GG (2004b) The effect of vegetation density on canopy sub-layer turbulence. Boundary-Layer Meteorol 111(3):565–587

    Article  Google Scholar 

  • Polyanin AD, Zaitsev VF (2003) Handbook of exact solutions for ordinary differential equations. CRC Press, Boca Raton, 720pp

  • Raupach MR, Thom AS (1981) Turbulence in and above plant canopies. Annu Rev Fluid Mech 13:97–129

    Article  Google Scholar 

  • Raupach MR, Finnigan JJ, Brunei Y (1996) Coherent eddies and turbulence in vegetation canopies: the mixing-layer analogy. Boundary-Layer Meteorol 78(3):351–382

    Article  Google Scholar 

  • Ross AN (2008) Large-eddy simulations of flow over forested ridges. Boundary-Layer Meteorol 128(1):59–76

    Article  Google Scholar 

  • Shaw RH, Patton EG (2003) Canopy element influences on resolved- and subgrid-scale energy within a large-eddy simulation. Agric For Meteorol 115(1–2):5–17

    Article  Google Scholar 

  • Shaw RH, Schumann U (1992) Large-eddy simulation of turbulent-flow above and within a forest. Boundary-Layer Meteorol 61(1–2):47–64

    Article  Google Scholar 

  • Shaw RH, Finnigan JJ, Patton EG, Fitzmaurice L (2006) Eddy structure near the plant canopy interface. Preprints, 27th conference on agricultural and forest meteorology and the 17th symposium boundary layers and turbulence, San Diego, USA, May 2006, Paper J2.1

  • Stevens B (2000) Quasi-steady analysis of a pbl model with an eddy-diffusivity profile and nonlocal fluxes. Mon Weather Rev 128:824–836

    Article  Google Scholar 

  • Su HB, Schmid HP, Vogel CS, Curtis PS (2008) Effects of canopy morphology and thermal stability on mean flow and turbulence statistics observed inside a mixed hardwood forest. Agric For Meteorol 148(6–7): 862–882

    Article  Google Scholar 

  • Thom AS (1971) Momentum absorption by vegetation. Q J R Meteorol Soc 97:414–428

    Article  Google Scholar 

  • Wang W (2009) The influence of thermally-induced mesoscale circulations on turbulence statistics over an idealized urban area under a zero background wind. Boundary-Layer Meteorology 131(3):403–423

    Article  Google Scholar 

  • Wang W (2010) The influence of topography on single-tower-based carbon flux measurements under unstable conditions: A modeling perspective. Theor Appl Climatol 99(1):125–138

    Article  Google Scholar 

  • Wang W (2012) An analytical model for mean wind profiles in sparse canopies. Boundary-Layer Meteorol 142(3):383–399

    Article  Google Scholar 

  • Wang W, Davis KJ (2008) A numerical study of the influence of a clearcut on eddy-covariance fluxes of CO\(_{2}\) measured above a forest. Agric For Meteorol 148(10):1488–1500

    Google Scholar 

  • Wang W, Rotach M (2010) Flux footprints over an undulating surface. Boundary-Layer Meteorol 136(2): 325–340

    Article  Google Scholar 

  • Wang W, Yi C (2013) A new nonlinear analytical model for canopy flow over a forested hill. Theor Appl Climatol 109:549–563

    Google Scholar 

  • Wilson NR, Shaw RH (1977) A higher order closure model for canopy flow. J Appl Meteorol 16:1198–1205

    Article  Google Scholar 

  • Yi C (2008) Momentum transfer within canopies. J Appl Meteorol Climatol 47:262–275

    Article  Google Scholar 

  • Zeng P, Takahashi H (2000) A first-order closure model for the wind flow within and above vegetation canopies. Agric For Meteorol 103:301–313

    Article  Google Scholar 

Download references

Acknowledgments

The author is grateful to two anonymous reviewers for their constructive comments and suggestions. This work is partly supported by the National Natural Science Foundation of China under Grant No. 41075039 and 41375058.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Weiguo Wang.

Appendices

Appendices

1.1 A.1 LES Experiments

LES experiments were conducted to derive the dependence on canopy density of wind and stress profiles within a canopy. The model described in Bryan and Fritsch (2002) was used to simulate canopy flow over a flat terrain under neutral conditions; this model has been successfully applied to simulating atmospheric boundary-layer turbulence under various conditions (Kang and Davis 2008; Wang and Davis 2008; Wang 2009, 2010; Wang and Rotach 2010). Canopies are assumed to be horizontally homogeneous and uniform in the vertical. To account for the drag imposed by canopy elements, a forcing term, \(D_i =-C_\mathrm{d} a_0 Vu_i\), is added to the momentum equation, where \(u_{i}\) is the resolved-scale instantaneous velocity, subscript \(i\) is taken to be 1, 2, or 3 referring to the velocity or force in the \(x\), \(y\), or \(z\) directions, respectively. \(V\) represents the resolved-scale instantaneous wind speed, while \(C_\mathrm{d} = 0.25\). A turbulence kinetic energy based subgrid-scale model (Deardorff 1980) is used to parametrize the subgrid-scale turbulence, with an inclusion of the effects of the canopy drag on kinetic energy, such as wake effects (Shaw and Schumann 1992; Dwyer et al. 1997; Shaw and Patton 2003; Patton and Katul 2009). A third-order Runge–Kutta time integration scheme is used, with a third-order vertical advection scheme and fifth-order horizontal advection scheme. The simulation domain is \(660 \times 330\;\hbox {m}^{2}\) in the streamwise \((x)\) and spanwise \((y)\) directions, respectively, with a horizontal grid of 1 m. The vertical grid size is 0.5 m below 16 and 3.5 m above 50 m, and varies linearly in between, with the top of the domain at 330 m. Given the limited computational resource, the domain is large enough to capture most of the scales of motions in the canopy boundary layer, according to Finnigan et al. (2009) and Patton and Katul (2009). A logarithmic law is used between the ground and the first model level, with a roughness length of 0.05 m. The lateral boundary condition is cyclic, and a rigid lid is used as the top boundary condition. Turbulence is triggered by randomly adding perturbations to initial velocity fields, while the Coriolis force is not considered for this small-scale simulation. Flow is driven by a pressure gradient of \(3.2\times 10^{-5}\) Pa m\(^{-1}\) imposed in the \(x\) direction above the canopy top (15 m), resulting in \(u_{*}\) being approximately 0.1 m \(\hbox {s}^{-1}\) after the turbulent flow reaches an equilibrium state. The model is integrated for 15 h for each run, and results from the last 5 h are averaged over time and in the \(x\) and \(y\) directions to approximate the ensemble averages of atmospheric variables, such as the first and second moments of the turbulent velocity fields. Seven LES runs with \(FAI = 0.1, 0.2, 0.4, 0.8, 1.2, 3.0\), and 5.0, ranging from sparse to dense canopies, were made.

1.2 A.2 Turbulence Closure

Ayotte et al. (1999) showed that standard parametrizations (without canopies) for the budget equations of the second moments may be appropriate for canopy flow if the energy dissipation is adjusted to reflect canopy dynamics (Finnigan 2000). Starting from the budget equation for the shearing stress in the time- and volume-averaged flow, Finnigan and Belcher (2004) derived a simple non-local closure parametrization scheme for the stress in the canopy roughness sublayer. For a steady horizontally homogeneous flow, the parametrized budget equation for the shearing stress can be written, after some rearrangement and simplifications, as

$$\begin{aligned} -\overline{{u^{\prime }}{w^{\prime }}} =K(z)\frac{\partial U}{\partial z}- \frac{\partial }{\partial z}\left( {A(z)\frac{\partial \overline{{u^{\prime }}{w^{\prime }}}}{\partial z}}\right) , \end{aligned}$$
(20)

where we have used the same symbols as in Finnigan and Belcher (2004), the dimension of coefficient \(A(z)\) is the squared length, \(A(z)\propto l^{2}\), where l is a height-dependent length scale. The first term on the right-hand side is the product of an eddy diffusivity and the local vertical gradient of mean wind speed and is termed a local term. The second term is associated with the parametrized turbulence/pressure transport and pressure-strain processes, which may be called a non-local term. If the non-local term is negligible, Eq. 20 can be approximated as Eq. 4. To use Eq. 20 in the analytical model, we have to simplify the second term. Here a scaling method is used, and by expanding the non-local term, we have,

$$\begin{aligned} \frac{\partial }{\partial z}\left( {A(z)\frac{\partial \overline{{u^{\prime }}{w^{\prime }}}}{\partial z}}\right) =\frac{\partial A(z)}{\partial z}\frac{\partial \overline{{u^{\prime }}{w^{\prime }}}}{\partial z}\;+A(z)\frac{\partial ^{2}\overline{{u^{\prime }}{w^{\prime }}}}{\partial z^{2}}. \end{aligned}$$
(21)

Finnigan and Belcher (2004) postulated that the length scale is a constant for a dense canopy, but for sparse and open canopies, the ground stress is not negligible, and therefore we assume that the length scale to first order is proportional to height above the ground, i.e., \(A(z)\propto z^{2}\). As a result, with \(A(z)\propto z^{2}\), the stress (\(\overline{{u^{\prime }}{w^{\prime }}})\) scaled by \(u_*^2\) and height by h, the first term on the right-hand side of Eq. 21 can be scaled with \(u_*^2 z/h\), and the second term can be scaled with \(u_*^2 \left( {z/h}\right) ^{2}\), which is smaller than \(u_*^2 z/h\) for \(z<\) h within the canopy. Therefore, with the second derivative term being neglected, the non-local term can be scaled with \(u_*^2 z/h\). To a first-order approximation, we can estimate the non-local term to be proportional to \(u_*^2 z/h\), i.e. \({f}_{n}u_*^2 z/h\), where \({f}_{n}\) is a height-independent coefficient. Despite the simplification, the parametrized stress (Eq. 9) can reasonably reproduce the stress profiles within canopies, with errors being within 10–20 % compared with laboratory measurements (Fig. 4) and LES results (Fig. 1).

1.3 A.3 Wind Profile as FAI Approaches Zero

Here we show that Eq. 11 reduces to the standard logarithmic wind profile in the case of canopy density being zero. As the order (\(\nu \)) is fixed and \(x\rightarrow 0\), the limiting forms of the modified Bessel functions are,

$$\begin{aligned} I_{\nu } (x)&\rightarrow \frac{1}{2}\frac{x^{\nu }}{\Gamma (\nu +1)},\end{aligned}$$
(22)
$$\begin{aligned} K_0 (x)&\rightarrow -\ln (x),\end{aligned}$$
(23)
$$\begin{aligned} K_1 (x)&\rightarrow \frac{\Gamma (1)}{x}, \end{aligned}$$
(24)

where \(\Gamma (n)=(n-1)!\) is the gamma function. As FAI decreases to zero, \(f_n\) approaches zero and \(s_h\) goes to 1. In this case, \(g(h)\) tends to zero. With the above properties, and neglecting the pressure gradient forcing term, the integration constants (Eq. 16) become,

$$\begin{aligned} C_1&\rightarrow -\frac{4u_*}{\kappa }\ln (g(z_0)),\end{aligned}$$
(25)
$$\begin{aligned} C_2&\rightarrow -\frac{2u_*}{\kappa },\end{aligned}$$
(26)
$$\begin{aligned} I_0 (g(z))&\rightarrow \frac{1}{2}, \end{aligned}$$
(27)

and

$$\begin{aligned} K_0 (g(z))\rightarrow -\ln (g(z)). \end{aligned}$$
(28)

Therefore, substituting Eqs. 2528 into Eq. 11, leads to,

$$\begin{aligned} U(z)&= C_1 I_0 (g(z))+C_2 K_0 (g(z))\nonumber \\&\rightarrow -\frac{2u_*}{\kappa }\ln (g(z_0))+\frac{2u_*}{\kappa }\ln (g(z))=\frac{u_*}{k}\ln \left( {\frac{z}{z_0}}\right) . \end{aligned}$$
(29)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Wang, W. Analytically Modelling Mean Wind and Stress Profiles in Canopies. Boundary-Layer Meteorol 151, 239–256 (2014). https://doi.org/10.1007/s10546-013-9899-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10546-013-9899-6

Keywords

Navigation