Roughness Sub Layers John Finnigan, Roger Shaw, Ned Patton, Ian Harman

Similar documents
The Urban Canopy and the Plant Canopy

J2.1 EDDY STRUCTURE NEAR THE PLANT CANOPY INTERFACE

The plant canopy layer and roughness sublayer. 16 Oct., 2012

The structure of canopy turbulence and its implication to scalar dispersion

The Atmospheric Boundary Layer. The Surface Energy Balance (9.2)

LES, lab and empirical representations of a neutral flow over canopy

Wind and turbulence experience strong gradients in vegetation. How do we deal with this? We have to predict wind and turbulence profiles through the

First order turbulence closure for modelling complex canopy flows

Turbulent boundary layer

TURBULENT STATISTICS OF NEUTRALLY STRATIFIED FLOW WITHIN AND ABOVE A SPARSE FOREST FROM LARGE-EDDY SIMULATION AND FIELD OBSERVATIONS

Temperature fronts and vortical structures in turbulent stably stratified atmospheric boundary layers

Keywords: Large-eddy simulation, Turbulent coherent structure, Four quadrant analysis, Integral scale

Logarithmic velocity profile in the atmospheric (rough wall) boundary layer

Canopy flow. Acknowledged contribution from Corey Markfort, Iowa State U

ESPM 129 Biometeorology Wind and Turbulence, Part 2, Canopy Air Space: Observations and Principles

A NOTE ON THE CONTRIBUTION OF DISPERSIVE FLUXES TO MOMENTUM TRANSFER WITHIN CANOPIES. Research Note

Note the diverse scales of eddy motion and self-similar appearance at different lengthscales of the turbulence in this water jet. Only eddies of size

DNS STUDY OF TURBULENT HEAT TRANSFER IN A SPANWISE ROTATING SQUARE DUCT

Lecture 2. Turbulent Flow

Footprints: outline Üllar Rannik University of Helsinki

BOUNDARY LAYER STRUCTURE SPECIFICATION

This is the first of several lectures on flux measurements. We will start with the simplest and earliest method, flux gradient or K theory techniques

The role of coherent structures in subfilter-scale dissipation of turbulence measured in the atmospheric surface layer

Dimensionality influence on energy, enstrophy and passive scalar transport.

(Wind profile) Chapter five. 5.1 The Nature of Airflow over the surface:

Donald Slinn, Murray D. Levine

τ xz = τ measured close to the the surface (often at z=5m) these three scales represent inner unit or near wall normalization


First-order turbulence closure for modelling complex canopy flows

Part I: Overview of modeling concepts and techniques Part II: Modeling neutrally stratified boundary layer flows

Atm S 547 Boundary Layer Meteorology

TWO-POINT CORRELATION ANALYSIS OF NEUTRALLY STRATIFIED FLOW WITHIN AND ABOVE A FOREST FROM LARGE-EDDY SIMULATION

Prototype Instabilities

Chuichi Arakawa Graduate School of Interdisciplinary Information Studies, the University of Tokyo. Chuichi Arakawa

6 VORTICITY DYNAMICS 41

Journal of Fluid Science and Technology

LECTURE 28. The Planetary Boundary Layer

Turbulence is a ubiquitous phenomenon in environmental fluid mechanics that dramatically affects flow structure and mixing.

Fluid Mechanics. Chapter 9 Surface Resistance. Dr. Amer Khalil Ababneh

COMMENTS ON "FLUX-GRADIENT RELATIONSHIP, SELF-CORRELATION AND INTERMITTENCY IN THE STABLE BOUNDARY LAYER" Zbigniew Sorbjan

Turbulent momentum flux characterization using extended multiresolution analysis

DESTABILISING EFFECTS OF PLANT FLEXIBILITY IN AIR AND AQUATIC VEGETATION CANOPY FLOWS

Principles of Convection

Homogeneous Turbulence Dynamics

HORSEHOE VORTICES IN UNIFORMLY SHEARED TURBULENCE

Part I: Dry Convection

2013 Annual Report for Project on Isopycnal Transport and Mixing of Tracers by Submesoscale Flows Formed at Wind-Driven Ocean Fronts

Chapter 5 Phenomena of laminar-turbulent boundary layer transition (including free shear layers)

Before we consider two canonical turbulent flows we need a general description of turbulence.

PHYS 432 Physics of Fluids: Instabilities

6.4 EXPERIMENTAL DETERMINATION OF THE TURBULENT KINETIC ENERGY BUDGET WITHIN AND ABOVE AN URBAN CANOPY

Large Eddy Simulation of an Inhomogeneous Atmospheric Boundary Layer under Neutral Conditions

Transition to turbulence in plane Poiseuille flow

Multichoice (22 x 1 2 % = 11%)

For example, for values of A x = 0 m /s, f 0 s, and L = 0 km, then E h = 0. and the motion may be influenced by horizontal friction if Corioli

A model for scalar advection inside canopies and application to footprint investigation

The Stable Boundary layer

DAY 19: Boundary Layer

meters, we can re-arrange this expression to give

2.3 The Turbulent Flat Plate Boundary Layer

Chapter (3) TURBULENCE KINETIC ENERGY

Lecture 3: The Navier-Stokes Equations: Topological aspects

Turbulence. 2. Reynolds number is an indicator for turbulence in a fluid stream

A NUMERICAL STUDY OF NOCTURNAL WAVELIKE MOTION IN FORESTS. 1. Introduction

Boundary-Layer Theory

On the turbulent Prandtl number in homogeneous stably stratified turbulence

General introduction to Hydrodynamic Instabilities

Analysis and modelling of the effective reaction rate in a developing mixing layer using OpenFOAM R libraries

Contents. I Introduction 1. Preface. xiii

THE EFFECT OF STRATIFICATION ON THE ROUGHNESS LENGTH AN DISPLACEMENT HEIGHT

Turbulent flow over anisotropic porous media

Symmetry of Turbulent Characteristics Inside Urban Intersection

7. Basics of Turbulent Flow Figure 1.

Spring Semester 2011 March 1, 2011

Chapter 4. Gravity Waves in Shear. 4.1 Non-rotating shear flow

EXCITATION OF GÖRTLER-INSTABILITY MODES IN CONCAVE-WALL BOUNDARY LAYER BY LONGITUDINAL FREESTREAM VORTICES

Frictional boundary layers

Road to Turbulence in Stratified Flow

Boundary Layer Meteorology The wind that shakes the buildings

A Discussion on The Effect of Mesh Resolution on Convective Boundary Layer Statistics and Structures Generated by Large-Eddy Simulation by Sullivan

Direct numerical simulation of salt sheets and turbulence in a double-diffusive shear layer

Colloquium FLUID DYNAMICS 2012 Institute of Thermomechanics AS CR, v.v.i., Prague, October 24-26, 2012 p.

UC San Diego International Symposium on Stratified Flows

April 1990 T. Watanabe and J. Kondo 227. The Influence of Canopy Structure and Density. upon the Mixing Length within and above Vegetation

O. A Survey of Critical Experiments

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

Introduction to Turbulence AEEM Why study turbulent flows?

Physics of irregularity flow and tumultuous conversion in shear flows

Atmospheric stability parameters and sea storm severity

arxiv: v1 [physics.flu-dyn] 7 Dec 2018

Atmospheric Boundary Layers

UNIT II CONVECTION HEAT TRANSFER

1. Introduction, tensors, kinematics

Direct and Large Eddy Simulation of stably stratified turbulent Ekman layers

LIDAR OBSERVATIONS OF FINE-SCALE ATMOSPHERIC GRAVITY WAVES IN THE NOCTURNAL BOUNDARY LAYER ABOVE AN ORCHARD CANOPY

ESS Turbulence and Diffusion in the Atmospheric Boundary-Layer : Winter 2017: Notes 1

Buoyancy Fluxes in a Stratified Fluid

NONLINEAR FEATURES IN EXPLICIT ALGEBRAIC MODELS FOR TURBULENT FLOWS WITH ACTIVE SCALARS

SMS 303: Integrative Marine

Surface layer parameterization in WRF

Transcription:

Roughness Sub Layers John Finnigan, Roger Shaw, Ned Patton, Ian Harman 1. Characteristics of the Roughness Sub layer With well understood caveats, the time averaged statistics of flow in the atmospheric surface layer over homogeneous surfaces are described by Monin-Obukhov Similarity Theory (MOST). However, it has been known for decades that MOST formulae fail close to rough surfaces like urban areas or vegetation canopies or surface waves. The failure is almost always in the sense that turbulent fluxes are higher than MOST would predict from the observed mean gradients. Three kinds of dynamical process can be invoked to account for these departures: Distortion of the mean flow and turbulence about the roughness elements Near field-far field diffusive effects for scalars Different mechanisms of turbulence generation We consider these in turn but first make a comment on statistics. Far from the surface we expect time-mean flow statistics to be independent of their x-y (horizontal) location. Near complex surfaces, the time-mean statistics become dependent on their x-y coordinates and we then form horizontally homogeneous statistical fields by averaging over thin horizontal slabs as well as in time. Mean flow distortion about the roughness elements Spatial correlations between time-mean variations around the spatial mean are termed Dispersive Fluxes and bear the same relationship to spatial averaging as Reynolds fluxes do to time averaging. Around smooth, regular spatial features like water waves or low topography, the dispersive fluxes can often be calculated deterministically. In such situations also, the effect of mean flow distortion on the Reynolds stresses can often be calculated accurately. These effects decay with height, affecting a layer whose depth is proportional to the horizontal scale of the surface heterogeneity. The affected layer can be thought of as a roughness sublayer but this is not what is commonly understood by the term. When the flow is over bluff elements, the resulting chaotic time-mean streamline pattern is called Lagrangian Turbulence and generally augments any temporal turbulence in mixing momentum and scalars. In this case, no general relationships between the Lagrangian and Reynolds turbulence can be stated. The region of mixing enhanced by Lagrangian turbulence can be taken as one definition of the roughness sub layer. Near field-far field diffusive effects for scalars A Lagrangian framework for modelling dispersion is appropriate for scalar transport in and above canopies. Raupach (1989) has shown that spatial perturbations to mean scalar gradients are dominated by nearby sources as turbulent dispersion in the near field of sources or sinks is slower than in the far field. The near field is delineated by the Lagrangian Integral scale which typically extends a distance of order the canopy height from the source. Conversely, the turbulent flux and background scalar concentration is dominated by distant sources as diffusion is faster in the far field. This near field-far field dichotomy decouples the local flux from the gradient and can cause deviations from MOST flux-gradient relationships just above the canopy. Within the canopy, in extreme cases it can lead to counter-gradient fluxes. It is a process that operates even when the turbulence is homogeneous and is important 1

whenever the Lagrangian integral scale is of significant size. The extent of near field influence, that is, the Lagrangian integral length scale is a second measure of RSL depth. Alternative mechanisms of turbulence production The mixing layer hypothesis (MLH) of Raupach et al (1996) proposed that the distinctive nature of turbulence over canopies or rough surfaces was the result of a generation mechanism that is fundamentally different from that operating in smooth wall boundary layers. It supposed that the inflected mean velocity profile that is ubiquitous at the rough surface-free air interface and which is inviscidly unstable, generated distinct energetic coherent eddies, whose scale was linked toδ ω, the characteristic scale of the inflected profile at the level of the canopy top, : δ = U h U hc. Velocity profiles without an inflection point are only h ( ) ( ) c ω c unstable if viscosity or buoyancy is present. A great deal of evidence collected over the last decade has confirmed the essentials of the MLH and established some limits to its applicability. In particular, although the horizontally averaged mean velocity profile over rough surfaces almost always displays an inflection point, the profile must be dynamically significant rather than an artefact of spatial averaging for MLH to be relevant. In this paper we develop the hypothesis that the roughness sublayer and its distinctive properties are the result of the MLH. Near field-far field effects and Lagrangian turbulence are complicating but not dominant features of the RSL while variations in the mean wind field caused by flow over smooth features like waves or gentle topography probably deserve a different title to distinguish their effects from those we are about to describe. Key differences between the RSL and the inertial sublayer (ISL) or Logarithmic Layer above it, where MOST obtains, are reflected in both the mean and conditional statistics: Turbulence statistics in the RSL/canopy layer scale on a single length while in the ISL, statistics scale on z-d (where d is the displacement height of the logarithmic profile). Empirical Orthogonal Function (EOF) spectra in the RSL/canopy converge rapidly (eg. 80% of the variance is captured by the first 5 eigenmodes), suggesting that the turbulence is dominated by distinct coherent structures of a single scale, whereas convergence in the ISL is slow (eg. 80% of the variance in the first 25-50 eigenmodes), denoting a wide range of eddy scales. The ratios of second moments in the RSL/canopy layer are more isotropic than in the ISL and the correlation coefficient r uw = uw σ u σ w is ~0.5 in the neutrally stratified RSL/canopy as compared to ~0.35 in the ISL. Putting this together with the observation that gradients are smaller in the RSL than in the ISL for the same flux suggests that RSL/canopy turbulence is in some sense more efficient at mixing. The turbulent Schmidt and Prandtl numbers are ~0.5 at the canopy- RSL interface compared with ~1.0 in the ISL. Transport of momentum and scalars in the RSL/canopy layer is dominated by the sweep quadrant whereas through the rest of the ISL, ejections dominate. 2

2. Turbulence Structure in the Roughness Sub layer A model of the turbulent eddy structure over smooth walls in boundary layers and channel flows has been developed by Adrian and various collaborators using sophisticated conditional sampling techniques. Flow in an ISL above a rough wall is very similar (although not identical) to that over a smooth wall and we will assume that Adrian s model is relevant there also. His model envisages the synergistic development of trains of head-up (HU) hairpin or horseshoe vortices aligned around cores that consist of a low-speed flow region, elongated in the streamwise direction. The sense of rotation of the HU hairpins is as if the spanwise vortex lines of the mean flow have been deflected upwards and then stretched and rotated by the mean shear. As a result, an ejection is generated between the legs of each HU hairpin and these ejections merge in the low speed core. Sweeps are weak or non existent. We have applied analogous conditional sampling approaches as well as EOF analysis to field data, wind tunnel simulations and LES models of uniform canopy flow. By interpreting our results in the light of the MLH theory we have arrived at a consistent picture of RSL turbulent structure that accounts for its differences from that in the ISL. The model has the following elements (Shaw et al., 2006): The hydrodynamic instability of the inflected velocity profile at canopy top produces Kelvin-Helmholtz waves with regions of spanwise vorticity of alternating sign and streamwise spacing proportional toδ ω. This perturbed vorticity field is itself unstable and the vorticity clumps into coherent spanwise Stuart vortices that retain the streamwise spacing of the Kelvin-Helmholtz waves. These Stuart vortices can be distorted by ambient turbulent eddies to form head up hairpins (deflection away from the wall) or head down (HD) hairpins (deflection towards the wall). The porous canopy layer allows strong deflections towards the wall that would be blocked near a solid surface. Secondary instabilities of this train of vortices provide strong selection for the spanwise scale of the hairpins. They also suggest that HU and HD hairpins will be formed in pairs, a prediction confirmed by our measurements. Similarly, the canopy depth h c limits the horizontal scale of strong gusts directed towards the wall: they cannot have horizontal extent much larger than h c. The strain and rotation experienced by a HD hairpin in a logarithmic mean velocity profile or a log-exponential profile, such as we find near the canopy top, is larger than that experienced by a HU so that with equal populations of HU and HD hairpins, HD s will be more energetic. HD hairpins generate sweeps between their legs so that near the canopy top sweeps dominate ejections. Further from the canopy and underlying surface, larger scale gusts can deflect the vortices but the blocking effect of the ground is now dominant so that strong upward deflections that produce HU hairpins and ejections dominate downward gusts that lead to sweeps. We have found empirically that the two hairpin types come in pairs with sweepproducing HD s overlying ejection-producing HU s. The strong convergence between the hairpins that this produces results in intense scalar microfronts. 3. A simple model of the Roughness Sub layer The fundamental role played by the inviscid instability in generating this distinct turbulent structure has prompted the construction of a simple RSL model based upon 3

its key scale, δ ω (Harman and Finnigan, 2006). This model views the RSL as a region influenced by the particular turbulence generated by the MLH. This influence decays exponentially above the canopy with an e-folding distance proportional toδ ω. Hence, unlike earlier models, we do not assume a precise depth for the RSL. The effects of the MLH turbulence are represented by a function $ φ (z δ ω ) that is analogous to the MOST φ ( zlmo )functions (where L mo is the Obukhov length) and combines multiplicatively with them. By matching a simple, one-parameter model of the canopy flow with the ISL, where φ ( zlmo )functions represent the effects of stability, through an RSL where both $ φ (z δ ω ) and φ ( zlmo ) affect the flux-gradient relationships, we find that the fundamental parameters of the log law, displacement height d and roughness length z 0, become strong functions of stability. When we consider the effect of this dependence on the parameterisation of momentum transfer to the surface in climate, mesoscale and boundary layer models, we find that that the effects are profound. The model has been applied to a range of forest canopy data varying in height, foliage density and heterogeneity and represents observations remarkably well. 4. Extension to scalars The simple canopy model used in the RSL model discussed above predicts an exponential decay of mean velocity into the canopy. For simple concentration boundary conditions on the foliage, the scalar analogue of this also predicts an exponential decay of mean scalar but with an exponent that is smaller than that for velocity by the factor( r Sc t ). Sc t ( Pr t ) is the turbulent Schmidt (Prandtl) number and In the upper canopy/rsl,sc t and Pr t are ~0.5 compared to ~1.0 in the ISL. The Stanton number is the ratio of the efficiency of scalar to momentum transport across canopy elements and r ~0.1so the factor r Sc t ~ 0.05. There are two immediate consequences of this. First, scalar gradients are much weaker than velocity gradients in the RSL with consequences for measurements of flux-gradient relationships and extrapolation of measurements in the RSL to the ISL and upper planetary boundary layer. Second, the difference in the intrinsic gradients of temperature and windspeed in the canopy ensures that for even moderate stability above the canopy, the local gradient Richardson Number within the canopy is large so that turbulence is strongly suppressed. This phenomenon effectively decouples the canopy flow from that above on stable nights and allows the generation of strong gravity currents even on very gentle slopes, if a canopy is present. 5. References Harman, I.N. and J.J. Finnigan, 2006. The impact of a dense canopy on the wind profile and evolution of a boundary layer. Proceedings of the American Meteorological Society 26th Conference on Boundary Layers and Turbulence, San Diego, May, 2006. Raupach MR. 1989. Stand Overstorey Processes. Philos. Trans. R. Soc. Lond. B 4

324:175 90 Raupach, M.R., Finnigan, J.J., and Brunet, Y. (1996). Coherent eddies in vegetation canopies - the mixing layer analogy. Boundary-Layer Meteorol., 78, 351-382. Shaw, R.H., J.J. Finnigan, E.G. Patton and Fitzmaurice, L., (2006) Eddy Structure near the plant canopy interface. Proceedings of the American Meteorological Society 26th Conference on Boundary Layers and Turbulence, San Diego, May, 2006. 5