pressure distributions on a cube in a simultated thunderstorm downburts part a

Upload: juan-carlos-gonzalez

Post on 04-Apr-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    1/22

    Journal of Wind Engineering

    and Industrial Aerodynamics 90 (2002) 711732

    Pressure distributions on a cube in a

    simulated thunderstorm downburstPart A:

    stationary downburst observations

    M.T. Chay, C.W. Letchford*Department of Civil Engineering Wind Science and Engineering Center, Texas Tech University,

    Lubbock, TX 79409-1023, USA

    Received 10 December 2001; received in revised form 19 March 2002; accepted 20 March 2002

    Abstract

    Thunderstorms are responsible for a large amount of wind-induced damage around the

    world. It is known that the wind characteristics in thunderstorms, particularly downbursts,

    differ significantly from those of synoptic scale boundary layer winds. This paper describes a

    study aimed at simulating the flow structure in a downburst and obtaining the pressure field on

    a cube immersed in such a flow. Part A presents the data obtained from a stationary wall jet

    simulation of a thunderstorm downburst, while Part B presents the data from a moving

    downburst simulation. The pressure distributions on the cube are compared with data from

    uniform and boundary layer flows. r 2002 Elsevier Science Ltd. All rights reserved.

    1. Introduction

    An atmospheric boundary layer wind profile currently forms the basis for

    calculating wind loads on structures [13], and has been the profile employed in wind

    tunnel simulations to obtain wind loading data either explicitly for specific buildings

    or implicitly through codified data for generic building shapes. However, thunder-

    storms are responsible for design wind speeds in many parts of the world [4,5] and

    the wind characteristics of thunderstorms are known to be different from boundary

    layer wind profiles. From these differences it may be anticipated that wind loading of

    both low- and high-rise buildings will be significantly different from those in

    traditional boundary layer flows. In particular, at low level the wind profile is more

    *Corresponding author. Tel.: +1-806-742-3476; fax: +1-806-742-3446.

    E-mail address: [email protected] (C.W. Letchford).

    0167-6105/02/$ - see front matter r 2002 Elsevier Science Ltd. All rights reserved.

    PII: S 0 1 6 7 - 6 1 0 5 ( 0 2 ) 0 0 1 5 8 - 7

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    2/22

    uniform with height, potentially leading to increased loads on low-rise structures,

    while at high level the reduction in wind velocity with elevation could reduce the

    wind loads on high-rise buildings. Full-scale observations [69], and physical

    simulations [1012] using wall jets have helped give a better understanding of thesephenomena and Letchford et al. [5] recently reviewed this work, however, there has

    been little attempt to assess and quantify the impact of thunderstorm wind profiles

    on the wind loading of structures.

    This study aimed to establish a viable simulation of a thunderstorm downburst

    and obtain surface pressure distributions on a generic building shapea cube

    immersed in this flow. Comparisons would then be made with pressure distributions

    in other flow simulations, namely uniform and turbulent boundary layer flows to

    ascertain the significance of the new impinging flow type. The downburst was

    simulated by a stationary wall jet, which has previously been shown [1012] to give a

    reasonable representation of the mean velocity profile (mean here applies to the

    short time averaged (several minutes) wind velocities measured at full-scale). To

    obtain better kinematic similarity of the flow and hence report meaningful unsteady

    pressure measurements, a moving wall jet was constructed so that the transient

    characteristics of a thunderstorm gust front could be obtained. Part A of this paper

    presents the results from the stationary wall jet simulation of the downburst, while

    Part B [24] presents results from the moving simulation. In the following section the

    basic characteristics of thunderstorm downbursts are reviewed. Section 3 reviews

    previous wind pressure measurements on cubes. In Section 4, the wall jet

    and velocity characteristics of the stationary downburst simulation are described.Section 5 compares the mean pressure distributions on a cube immersed in the

    stationary wall jet with earlier studies in uniform and boundary layer flows. A

    discussion of the experimental results is presented in Section 6.

    2. Thunderstorm downbursts characteristics

    Letchford et al. [5] discuss the general characteristics of thunderstorms.

    Fundamentally, convection drives an updraft, which transports warm moist, more

    buoyant, air to great elevations. Subsequently, the moisture in this air condenses,cools, and the upward motion is halted. The now colder more dense air begins to

    accelerate toward the ground as a downdraft. Downbursts occur when a strong

    downdraft collides with the surface of the earth and diverges. Close to the point of

    impact the flow resembles that of a wall jet. As the flow spreads out over the ground

    it behaves as a gravity or density current. The flow field created by such an event,

    particularly near the impact point, varies from an atmospheric boundary layer wind

    field in a number of fundamental ways.

    Firstly, the traditional boundary layer profile, of increasing wind velocity with

    height is no longer valid as a region of accelerated flow exists close to the surface

    with a decrease in velocity with height. Downdrafts and consequently downburstsexist over a range of scales; Fujita [6] classified them as microbursts if the area of

    damaging winds was o4 km in extent and macrobursts if >4 km. Typically

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732712

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    3/22

    microbursts produced the higher winds. The full-scale observations of downbursts

    during Fujitas NIMROD experiments [6] and the JAWS [8] experiments produced

    quantitatively similar results for microbursts. The Hjelmfelt [8] summary of the

    JAWS results is reproduced in Fig. 1. On average, the maximum wind velocity

    occurred at a height ofE80 m at a distance ofE1.5 km from the point of impact.

    This was for an average downburst diameter of 1.8 km. Thus, the strongest winds

    were observed within about one downdraft diameter from its point of impact. The

    primary data sources for wind velocities were three Doppler radars [8] and, a lowlevel, automated mesonets with an average spacing of 4 km. The Doppler radars

    were able to scan approximately every 2.5 min. Wilson et al. [9] present a full

    discussion on the full-scale analysis techniques and estimation of errors.

    Secondly, downdrafts often retain large amounts of the translational momentum

    of the parent storm. Storm velocities can be as much as a third the velocity of the

    downdraft [5]. The lateral motion of the downdraft causes an increase in the peak

    downburst velocity in front of the storm, and a decrease in the velocity on the

    trailing side. Due to the translating motion of the downdraft, stationary objects

    experience downburst winds as non-stationary events. The effect of the transient

    nature of such phenomena, in particular, the vortex ring at the leading edge of thedowndraft has not been previously investigated and is the subject of the Part B of

    this paper.

    Fig. 1. Velocity profile of a typical microburst during JAWS (Hjelmfelt [8]).

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732 713

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    4/22

    Thirdly, current procedures for assessing wind loads on structures assume aconstant ambient atmospheric pressure. However, significant variation of atmo-

    spheric pressure can be experienced within the flow field of a downburst. Stagnation

    occurs in the central region beneath the downdraft as it approaches the ground,

    forming a high-pressure dome known as a mesohigh (Fig. 2). A low-pressure ring

    forms as the downburst flow diverges and accelerates to the peak horizontal velocity.

    The relative magnitude of the pressure decrease is dependant on the translational

    velocity of the storm [7]. Beyond the low-pressure ring, deceleration of the outflow

    forms another slightly less intense high-pressure region, outside of which the pressure

    returns to the ambient pressure. The varying pressure field of a downburst may have

    serious implications with respect to design loads on structures. Fujita [7] speculatedthat these rapid pressure changes could be as high as 23 hPa, which may result in

    a significant increase in the load applied to sealed structures in the outflow region. As

    the downburst is a transient phenomenon, the pressure variation described by Fujita

    is naturally a transient one as well and is due to the accelerations within the flow.

    3. Previous investigations of pressures on a cube

    Numerous previous studies [1320] provide a comprehensive description of the

    pressure distribution over a cube immersed in both uniform (i.e., a wind fieldshowing little or no change in velocity as a function of height) and boundary layer

    wind fields (i.e., a wind field showing an increase in velocity as a function of height

    Fig. 2. The pressure field of a microburst (Fujita [7]).

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732714

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    5/22

    due to the effects of surface roughness). These studies have involved physical and

    numerical simulations, and full-scale measurements. The following paragraphs detail

    a few of the more notable papers.

    Baines [13] undertook early wind tunnel studies on a cube immersed in uniformand boundary layer flows. Unfortunately, no turbulence characteristics of the

    boundary layer flow are reported. He reports pressure distributions on each face of

    the cube, for the cube positioned with one face normal to the direction of flow.

    Castro and Robins [14] also undertook wind tunnel tests on a cube. Pressures were

    measured on a 60 mm cube in a uniform flow, and a 200 mm cube in a 2 m high

    boundary layer (constituting a scale of between 1:1000 and 1:300 for the boundary

    layer case) with a turbulence intensity of 27% at eaves height. Measurements were

    made with the cube face normal to the flow and at 451 to the flow. Castro and

    Robins also investigated the effects of varying the boundary layer characteristics

    acting on the cube. They found that a higher turbulence intensity favored

    reattachment of flow over the cube, and hence a reduced suction in the reattached

    region of the roof and leeward wall, for a flow normal to one face of the model.

    The study reported by H .olscher and Niemann [16], involved a comparative

    experiment of pressures over a cube across some 15 wind tunnels in Europe. There is

    a surprising amount of scatter in the results much of which was attributed to

    differences in turbulence intensities in the various boundary layer simulations.

    Paterson and Apelt [18] performed a numerical simulation using a k e

    turbulence model to simulate a boundary layer flow around a cube with one face

    normal to the direction of flow. They performed the study under similar conditionsto those Castro and Robins [14] investigated. However, the Paterson and Apelt study

    yielded greater mean pressures near the windward edge of the roof than Castro and

    Robins. Paterson and Apelt observed that increasing turbulence intensity promoted

    reattachment on the roof of the cube, supporting Castro and Robins wind tunnel

    observations.

    Richards et al. [20] collected full-scale data using the Silsoe Cube, which stands

    6 m high and had 16 pressure transducers across its centerline. The cube is situated in

    an open field at the Silsoe Research Institute in the United Kingdom. Richards et al.

    observed pressures on the cube with winds perpendicular and at 451 to the centerline

    pressure tap distribution. Turbulence intensities between 12% and 19% at cubeeaves height occurred during the study.

    Fig. 3 shows the mean pressure coefficient distribution along the developed

    centerline of the cube, with the windward face being between positions 0 and 1, the

    roof between 1 and 2 and the leeward wall between 2 and 3. The flow is normal to

    the front face (01 orientation). The pressure coefficients were formed by using the

    velocity at the top of cube as the reference velocity for the dynamic pressure. Baines

    [13] and Castro and Robins [14] observed similar pressures along the centerline of a

    cube immersed in uniform flow for this flow direction. However, significant variation

    exists between the pressure distributions for boundary layer flow over a cube at this

    orientation. The size of the separated flow region, and consequently the magnitudeof the leeward wall pressure, represents the primary difference between the various

    studies and differences in turbulence intensities are the likely cause. Richards et al.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732 715

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    6/22

    [20] observed reattachment further downstream than the three boundary layer

    simulations, which is consistent with the reduced turbulence that occurred during the

    full-scale measurements.

    4. Downburst simulation

    4.1. The Moving Jet Wind Tunnel

    For the current study, an inverted wall jet capable of translational movement, and

    termed the Moving Jet Wind Tunnel, produced the downburst simulation. Fig. 4

    shows a schematic of the arrangement, which was a further development of that used

    by Letchford and Mans [21]. A 5.6 kW centrifugal blower running at 3450 rpm drove

    air through extensive flow conditioning in the form of a settling chamber, screens,

    two layers of honeycomb and a 4:1 contraction. The 0.51 m diameter jet (D) blew

    against an extensive flat test surface positioned 870 mm (1.7D) above its outlet. The

    test surface was smooth-painted plywood. The asymmetry caused by the adjacent

    wall was countered by leaving a 150 mm gap at the edge of the test surface. Full

    details may be found in Chay [22]. As downbursts typically contain much colder,denser, air than that surrounding, the effort here has been aimed at producing a non-

    entraining jet. Thus, the nozzle outlet should be close to the surface. As a further

    0

    1

    2

    3

    -1.5

    -1

    -0.5

    0

    0.5

    1

    1.5

    0 1 2

    Position

    Cp

    Castro and Robins, Uniform Flow

    Castro and Robins, Boundary Layer Flow

    Baines, Uniform Flow

    Baines, Boundary Layer Flow

    Paterson and Apelt, Boundary Layer Flow

    Richards et al., Full Scale Measurements

    3

    Fig. 3. Comparison of centerline pressures on a cube under various flow regimes with one face normal to

    the direction of flow.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732716

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    7/22

    means of reducing entrainment, a 50 mm lip around the nozzle outlet was added to

    retard shear layer development between the ambient air and the jet.

    The outlet velocity and turbulence intensity profiles at half a diameter above the

    jet outlet are shown in Figs. 5 and 6, respectively. Longitudinal refers to thetraversing direction of the moving jet while only half the lateral profile is shown, as

    this was the limit of the anemometer traverse mechanism. This location was chosen

    Switches

    Cube

    PitotZ

    X

    3.9m

    5m1.1m

    2.4m 0.15m

    Fig. 4. The Moving Jet Wind Tunnel.

    0

    2

    4

    6

    8

    10

    12

    -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8

    Diameters from Center

    Velocity(m/s)

    Longitudinal Traverse

    Lateral Traverse

    Fig. 5. Velocity profiles half a diameter above the jet outlet.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732 717

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    8/22

    to be well clear of the screen and honeycomb at the jet outlet and away from the

    influence of the ground plane. The average velocity here (Vref) was E10 m/s with a

    turbulence intensity ofE4%. Fig. 7 shows the velocity spectrum on the jet centerline

    half a diameter above the jet outlet and indicates little high frequency content in the

    jet and lower magnitude fluctuations than might be anticipated in a boundary layer

    flow. Fig. 8 shows the reduction in jet velocity as it approaches the testing surface.

    The decay is small above 1 jet diameter and decreases linearly below 0.7 jet diameters

    to zero at the surface.

    The blower was mounted on rails and could be translated manually atapproximately constant velocity of up to 2 m/s, as timed by a pair of switches

    mounted on the track 5 m apart. One switch was positioned directly under the model

    location (X 0) to synchronize jet motion and model pressures. The results for the

    moving jet are presented in Part B of this paper [24].

    A model cube of side length 30 mm was constructed from 2 mm thick Perspex and

    38 tappings of 1 mm diameter were located around the cube. Tappings were

    distributed along a vertical centerline with six equally spaced pressure taps located

    on the windward and leeward walls and seven equally spaced taps across the roof.

    An additional six equally spaced taps in a horizontal profile recorded pressures along

    one sidewall at mid-height. A dense grid of 13 taps provided a more detailed recordof pressures on one roof corner. The tappings were connected by 200 mm of 1.02 mm

    diameter tubing to a Scanivalve ZOC33 64Px pressure measurement system and

    0

    5

    10

    15

    20

    25

    30

    35

    40

    45

    50

    -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8

    Diameters from Center

    TurbulenceIntensity(%)

    Longatudinal Traverse

    Lateral Traverse

    Fig. 6. Turbulence intensity profiles half a diameter above the jet outlet.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732718

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    9/22

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    1.4

    1.6

    0.0 0.2 0.4 0.6 0.8 1.0 1.2

    V/Vref

    Z/D

    Fig. 8. Non-dimensional velocity decay profile between the jet and testing surface along the jet centerline.

    The outlet of the jet is located at Z=D 1:7:

    0.00001

    0.0001

    0.001

    0.01

    0.1

    1

    0.1 1 10 100

    Frequency (Hz)

    Magnitude((m/s)^2/Hz)

    Fig. 7. Velocity power spectrum at half a diameter above the center of the jet outlet.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732 719

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    10/22

    sampled at 400 Hz for 60 s. Dynamic calibration of the pressure measurement system

    revealed that there wasE10% amplitude magnification at 100 Hz. While corrections

    to the frequency response for stationary tests could have been undertaken, they

    would have been problematical for the transient moving jet tests (detailed in Part B)and for consistency and because the amplification was small, no corrections to

    pressure data were made. Furthermore, pressure spectra revealed negligible

    amplitude at frequencies above 100 Hz. The reference pressure was background

    atmospheric pressure in the laboratory, well away from the jet.

    A Cartesian coordinate system with an origin at the center of the base of the

    model, located on the centerline of the moving jet, was employed, with Z measured

    away from the surface and positive X being upstream of the model. The jet diameter

    (D) was used to non-dimensionalize all distance in this study.

    4.2. Experimental procedure

    Velocity profiles produced by the wall jet were obtained using a TSI IFA300 hot

    wire anemometer system and single wire hot film probes sampled at 200 Hz for 20 s

    and repeated several times. The hot wire was mounted parallel to the test surface.

    Velocities were non-dimensionalized by the centerline jet mean velocity (Vref)

    measured at the reference location of half a diameter beyond the jet outlet (at

    Z=D 1:2 and X=D 0) before and after each traverse. Vertical velocity profileswere obtained for various stationary positions of the jet away from the origin

    (0pX=Dp3) for heights ranging from 3 to 153 mm. At each X=D position of the jet,a mean velocity ratio between the reference location and at the roof or eaves height

    of the cube, 30 mm, was also obtained for later pressure coefficient reduction of the

    cube pressures.

    A miniature pitot-static tube positioned half a diameter above the jet outlet

    (Z=D 1:2), was sampled at the beginning and end of each pressure test run andwas used as an initial reference dynamic pressure when non-dimensionalizing cube

    pressures. Pressure measurements over the cube were obtained for stationary jet

    positions ranging from 0pX=Dp3 and for two cube orientations, 01 and 451 to aface. The jet-induced static pressure over the inverted ground surface was obtained

    by measuring the pressure at a flush-mounted tap at the model location (X=D 0)without the model in place. This was undertaken for the jet in positions 0pX=Dp3:

    4.3. Stationary jet velocity profiles

    Fig. 9 shows the mean velocity profiles over the test surface as a function of the

    position of the jet. The velocities have been non-dimensionalized by the outlet

    velocity of the jet at the reference location, while heights have been non-

    dimensionalized by the jet diameter. These velocities were obtained from the single

    film probe and as such represent the velocity magnitude. In regions close tostagnation and well above the test surface, the velocity will have significant vertical

    component. The height of the cube is also shown on the figure for reference.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732720

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    11/22

    When the jet is placed directly over the measuring point, X=D 0; the jet velocity

    decreases linearly to the stagnation point on the test surface. Note for this locationthe reference velocity is located at 1.2 diameters above the surface (i.e., at Z=D 1:2;V=Vref 1). The characteristic nose of a wall jet develops after 0.75 diameters,reaching maximum mean velocity at about 1 diameter and then gradually slowing

    and thickening as the test surface induces boundary layer growth. At X=D 1; thelargest velocity in the profile is approximately the same as the reference wind velocity

    (V=Vref 1). At this point in the flow, the wall jet is thin and the flow ispredominantly horizontal. The almost constant velocity with height indicates that

    the effect of surface roughness has not developed in this region, and that the

    boundary layer that must form over the testing surface is lower than the lowest

    measurement point (3 mm). Castro and Robins [13] observed for their uniform flowtests a boundary layer thickness of between 2 and 6 mm.

    The geometric scale of the current simulation may be estimated based on the

    velocity profiles of the stationary jet. Hjelmfelt [8] observed that the parent

    downdraft of a typical microburst had a 1.8 km diameter and that the maximum

    outflow winds occurred atE1.5 km from the center of the descending column of air.

    Based on these observations, the Moving Jet Wind Tunnel simulation has a

    geometric scale ofE1:35001:3000. However, as Hjelmfelt observed microbursts

    with diameters ranging between 1.2 and 3.1 km, the current simulation may represent

    a range of scales based upon these dimensions. A velocity scale is more difficult to

    determine as the full-scale velocities in the downdraft ranged from 6 to 22 m/s,compared with the B10 m/s here. It becomes more important to define a velocity

    scale when the jet translates and this is the subject of Part B [24].

    0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    0.35

    0 0.2 0.4 0.6 0.8 1 1.2 1.4

    V/Vref

    Z/D

    X/D=0.0

    X/D=0.5

    X/D=0.75

    X/D=1.0

    X/D=1.25

    X/D=1.5

    X/D=2.0

    X/D=3.0

    Fig. 9. Non-dimensional wind velocity profiles over the testing surface as a function of distance from the

    stagnation point X=D 0 for a stationary jet.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732 721

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    12/22

    The turbulence intensities (standard deviation/mean velocity) for the wall jet as a

    function of distance from jet stagnation point (X=D 0) are presented in Fig. 10.Little turbulence occurred in the flow close to the stagnation point, although as themean velocity tended to zero, the turbulence intensity became very large. When

    positioned between X=D 0:5 and 0:75; the jet produced turbulence intensities ofE13% over the height of the cube. At the point of maximum mean wind velocity,

    X=D 1; the turbulence intensity was 20% over the height of the cube. At greaterdistances from stagnation the turbulence intensities become very large and possibly

    beyond the ability of the probe to accurately respond. This is because the mean

    velocity in these regions is tending to zero.

    Wood et al. [10] investigated the effect of varying the distance between the jet

    outlet and the testing surface (Zj) on the wind velocity profile. Fig. 11 comparesWoods results, with those of the current tests, and those of earlier tests at the

    University of Queensland [21] at the location of greatest velocity BX=D 1: Woodused a 300 mm diameter jet with Vref 20 m/s, and Letchford and Mans [21] a

    430 mm jet with Vref 7 m/s. It is clear that there is a distinct relationship between

    velocity profile and distance of jet from test surface (Zj). The closer the jet outlet to

    the test surface, the thicker and faster is the wall jet. This is not surprising as the

    simulation is effectively a free jet that will dissipate approximately as the square root

    of distance from the nozzle or outlet. In the studies shown in Fig. 11, the maximum

    mean velocities remained at distances of E1 diameter from jet stagnation,

    irrespective of distance between jet and wall (Zj).Hjelmfelt [8] analyzed the wind velocity profiles of eight full-scale downbursts.

    Fig. 12 shows a comparison of Hjelmfelts investigation to the wind velocity profiles

    0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    0.35

    0 20 40 60 80 100 120 140 160

    Turbulence Intensity (%)

    Z/D

    X/D=0.0

    X/D=0.5

    X/D=0.75

    X/D=1.0

    X/D=1.25

    X/D=1.5

    X/D=2.0

    X/D=3.0

    Eaves Height

    Fig. 10. Turbulence intensity profiles over the testing surface as a function of distance from the stagnation

    point X=D 0 for a stationary jet.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732722

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    13/22

    produced between 0:75pX=Dp1:5 from the current study. The velocities have been

    non-dimensionalized by the maximum velocity in each profile while heights havebeen non-dimensionalized by the elevation at which the maximum velocity occurred.

    The wind velocity profiles of the current study show less variation than the full-scale

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

    V/Vref

    Z/D

    Wood et al. Zj = 0.5D

    Letchford and Mans Zj = 1.25D

    Chay Zj = 1.7D

    Wood et al. Zj = 2.0D

    Wood et al. Zj = 5.0D

    Fig. 11. Comparison of non-dimensional velocity profiles at X=D 1 with various separations of jetoutlet from testing surface (Zj).

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    4.5

    0 0.2 0.4 0.6 0.8 1 1.2

    V/Vmax

    Z/Zmax

    full-scale minimum

    full-scale mean

    full-scale maximum

    X/D=0.75

    X/D = 1.0

    X/D=1.25

    X/D=1.5

    Fig. 12. Comparison of non-dimensional velocity profiles to full-scale data (Hjelmfelt [8]).

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732 723

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    14/22

    data, which is rather limited. However, the overall trends are similar, indicating that

    a wall jet is a reasonable representation of the mean wind profile.

    4.4. Static pressure field

    A single pressure tap on the testing surface recorded the static pressure variation

    within the simulated downburst. Eq. (1) shows the conversion of the mean static

    pressures at each jet location (X=D) to coefficient form with respect to the meandynamic pressure of the jet at the reference location. Ambient pressure in the

    laboratory away from the jet (PATMOS) provided the reference pressure for the

    transducers

    CP PSTATIC PATMOS

    12rV2ref

    : 1

    Fig. 13 shows the variation of static pressure beneath the wall jet as a function of

    distance from stagnation. The stationary jet produced a high-pressure region

    between 0pX=Dp0:25 approximately equal to the stagnation pressure at the outlet.This region represented the mesohigh of a downburst. The pressure coefficient at

    stagnation exceeded 1 as the jet produced a slightly non-uniform velocity profile over

    the cross-section of the outlet, increasing slightly in magnitude away from the center

    of the outlet, where the reference velocity was measured (Fig. 5). As the jet was

    positioned further from X=D 0:5 the static pressure showed a sharp decrease and

    quickly approached atmospheric pressure at X=D 1:5: The stationary jet did notcreate a negative static pressure region, as indicated by Fujita [7] in Fig. 2. This is

    because in this quasi-steady simulation, there are no transient ring vortices formed

    -

    0.2

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    00.511.522.533.5

    X/D

    Cp

    Fig. 13. The mean static pressure field of the stationary jet.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732724

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    15/22

    by the downdraft and which subsequently interact with the ground as suggested by

    Fujita [7]. This will be discussed in detail in Part B of this paper, which deals with the

    moving jet.

    5. Results

    5.1. Stationary jet pressure tests

    Pressures were measured over the cube for two orientations (01 and 451), with the

    jet located in the range 0pX=Dp3: Only 01 results are discussed in this paper.Eq. (2) indicates the conversion of mean surface pressures to coefficient form, with

    the ambient pressure in the laboratory away from the jet providing the reference

    pressure (PATMOS) and the mean dynamic pressure from the pitot-static tube at the

    reference location

    CPJ P PATMOS

    12rV2ref

    : 2

    Fig. 14 shows the variation of the mean pressure coefficient along the cube centerline

    as a function of jet position for wind perpendicular to one face (01). At the

    -1.5

    -1

    -0.5

    0

    0.5

    1

    1.5

    0 1 2

    Position

    Cpj

    X/D=0.0 X/D=0.5 X/D=0.75

    X/D=1.0 X/D=1.25 X/D=1.5

    X/D=2.0 X/D=3.0

    0

    1

    2

    3

    3

    Fig. 14. Mean pressure coefficients observed along the centerline of the cube.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732 725

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    16/22

    stagnation position, X=D 0; the jet produced pressure coefficients slightly higherthan 1 at all locations on the model. These are larger than 1 for the same reason that

    the surface static pressure was larger than 1 at stagnation, as discussed in Section 4.4

    above. Loading under these conditions would be particularly relevant to sealedstructures, where the internal pressure would remain at pre-downburst atmospheric

    pressure as represented by PATMOS in Eq. (2).

    As the jet is positioned further from the model, the influence of the jet stagnation

    pressure wanes while the effect of airflow becomes apparent with negative pressure

    coefficients developing in separated flow regions on the roof and lee wall for X=D >0:75: The mean pressure coefficient profile X=D 0:5 reflects the action of wind flowover the model, although the mean pressures on the faces of the cube are still

    positive.

    The mean pressure coefficient along the cube centerline showed very little

    variation over the windward face for most positions of the jet, while there was a

    gradual decrease in magnitude across the roof. The jet produced the greatest

    magnitude mean suction pressures when positioned in the range 1pX=Dp1:25; withthe dominant component due to wind flow as the static pressure becomes very small.

    The magnitude of the mean pressures decreased considerably as the jet moved

    beyond X=D 1:5:

    5.2. Comparison of mean pressure coefficients with earlier studies

    To facilitate comparison with earlier traditional wind tunnel studies, a rooftop oreaves height pressure coefficient was defined by Eq. (3). Here, use is made of the ratio

    of mean velocities between the reference velocity location and that at eaves height at

    X=D 1 to convert the coefficients defined earlier by Eq. (2). The X=D 1 locationwas chosen because it is where the velocities and pressures were largest, and therefore

    constitutes a design case

    CPE P PATMOS

    12rV2

    EAVES;X=D1

    : 3

    Fig. 15 presents a comparison between these pressure coefficients and Castro and

    Robins study [14] in uniform and turbulent boundary layer flows. The present

    centerline pressure coefficients bear closer resemblance to the uniform flow results,

    although they show slightly greater variation across the roof. On the front face the

    wall jet produces significantly higher positive pressures than measured by Castro and

    Robins for both their flow types. This apparent increase is due to the eaves height

    velocity being only some 90% of velocities at lower heights over the cube, as

    indicated in Fig. 9.

    Castro and Robins [14] and H .olscher and Niemann [16] identify turbulence

    intensity differences as a significant factor with respect to the pressure distributionover the roof of a cube and this could also have contributed to the variations

    observed here. At X=D 1; the stationary jet produced an approximately uniform

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732726

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    17/22

    velocity profile with a 20% turbulence intensity at eaves height, compared too0.5%

    in the uniform flow of Castro and Robins. Contrastingly, the eaves height turbulence

    intensity for Castro and Robins boundary layer flow wasB27%, indicating that the

    mean flow type plays a very significant role in the surface pressure distribution.

    5.3. Effect of the static pressure field

    The effect of the static pressure field produced by the wall jet on building pressures

    was examined by defining new pressure coefficients as indicated in Eq. (4), in which

    the raised static pressure of the wall jet (Fig. 13) is removed from the building

    pressures. Once again the mean eaves height dynamic pressure at X=D 1 was used

    CPS P PSTATIC;X=D12rV2

    EAVES;X=D1

    : 4

    Permeable structures, in which the internal pressure can quickly equilibrate to

    changes in atmospheric pressure, would likely experience the pressures defined by

    this coefficient.

    Fig. 16 shows these new pressure coefficients and it is seen that a significantreduction in the pressures generated on the cube at locations close to the center of the

    jet occurs. However, removal of the raised static pressure of the wall jet caused little

    -1.5

    -1

    -0.5

    0

    0.5

    1

    1.5

    0 1 2

    Position

    Cpe

    X/D=0.75

    X/D=1.0

    X/D=1.25

    Castro and Robins, Uniform Flow

    Castro and Robins, Boundary Layer Flow

    0

    1

    2

    3

    3

    Fig. 15. Comparison of centerline pressure coefficients of the stationary jet with earlier studies.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732 727

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    18/22

    change for pressure coefficients in the range 0:875pX=Dp1:25; which correspondsto the largest pressures/suctions under the action of the simulated downburst.

    Cassar [23], using a trial wall jet at CSIRO [9], undertook the only other study of

    pressures on a building model immersed in a wall jet known to the authors. This jet

    had an octagonal outlet ofE1.5 m 0.85 m, with an effective diameter (D) of

    1.05 m. Flow from the jet impinged on a smooth particleboard located 1.4 m away

    from the outlet (Zj=D 1:33). Fig. 17 shows a comparison of the mean wind velocity

    profiles at X=DB1 created by the jet used in the current study and the CSIRO jet.The velocities have been non-dimensionalized by the maximum velocity in the profile

    (12.7 m/s for CSIRO) while the heights have been non-dimensionalized by cube

    height, which was 100 mm for the CSIRO study. Clearly, the CSIRO model was

    relatively larger with the top of that cube extending into the shear layer region above

    the wall jet.

    Fig. 18 compares the centerline pressure coefficients over the cube for these two

    studies. The coefficients are defined by Eq. (4), excepting that for the present studys

    X=D 0:75 position, the eaves height dynamic pressure at X=D 0:75 has beenused to ensure compatibility with Cassars pressure coefficient definition. The shape

    of the mean pressure profiles is similar, with good agreement on the front face for theX=D 0:75 case, however, Cassars suction pressures are all much lower inmagnitude. It is likely that the differences in velocity profile in relation to the cube

    -1.5

    -1

    -0.5

    0

    0.5

    1

    1.5

    0 1 2

    Position

    Cps

    X/D=0.0 X/D=0.5 X/D=0.75

    X/D=1.0 X/D=1.25 X/D=1.5

    X/D=2.0 X/D=3.0

    0

    1

    2

    3

    3

    Fig. 16. Centerline pressures on the cube for 01 orientation, referenced against the static pressure of the

    diverging flow and expressed as a ratio of the eaves height dynamic pressure at X=D 1:

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732728

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    19/22

    -1.5

    -1

    -0.5

    0

    0.5

    1

    1.5

    0 1 2

    Position

    Cps

    Cassar (CSIRO Jet), X/D = 1.0

    Chay, X/D = 1.0

    Chay, X/D = 0.75

    0

    1

    2

    3

    3

    Fig. 18. Comparison of centerline pressure coefficient profiles created by the TTU Moving Jet Wind

    Tunnel and the CSIRO wall jet on a model cube.

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    4.5

    0 0.2 0.4 0.6 0.8 1 1.2

    V/Vmax

    Z/H

    Chay

    Cassar (CSIRO Jet)

    Fig. 17. Comparison of mean velocity profiles at X=D 1 for the TTU Moving Jet Wind Tunnel andCSIRO wall jet (Cube height=H).

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732 729

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    20/22

    height (Fig. 17), i.e., a cube three times larger in a flow nominally only twice as big,

    are responsible. In addition, the ambiguity in defining a representative static

    reference pressure for this type of flow could also explain the observed differences.

    6. Discussion

    This paper reports a study of mean pressures generated over a cube in a simulated

    thunderstorm downburst flow. A stationary wall jet was employed for this quasi-

    steady downburst simulation. Wall jets have been shown to have similar mean

    velocity characteristics of reported full-scale downbursts, in particular, the variation

    of velocity with height above the ground which leads to a velocity maximum at some

    50100 m at distances about 1 jet diameter from the stagnation point of the jet. Thegeometric scale of this simulation was estimated to be 1:3000.

    The velocity profiles produced by this simulation have been shown to relate to

    earlier studies and it has been shown that wall jet velocities close to stagnation are a

    function of separation distance between the jet outlet and the wall. The maximum

    mean velocity occurred at a distance of 1 jet diameter from the stagnation point and

    produced a nearly constant velocity distribution over the height of the cube.

    Turbulence intensities at this location were E20%.

    As expected the static surface pressure distribution over the wall produced by the

    jet varied significantly from maximum at stagnation to background atmospheric at a

    radius of about 1.5 jet diameters. Pressures generated over the cube could be dividedinto three regions based on location from jet stagnation (X=D 0):

    * Directly beneath the jet when almost all pressure is due to the static pressure field.* A transition region where pressures are due to both the static pressure field and

    diverging wall jet flow X=DB0:5:* A region in which almost all pressure experienced by the cube is due to the

    diverging flow of the wall jet X=D > 0:75

    Comparing the pressure distributions over the cube with conventional wind tunnel

    studies in both uniform and boundary layer flows indicated a greater likeness withuniform flow tests in the region of highest magnitude pressure around X=D 1: Thiswas attributed to the similarity of mean velocity profiles over the height of the cube

    in this region. However, the significantly greater turbulence in the wall jet lead to

    greater variation in separated flow regimes as might be expected given the

    importance of turbulence in separating shear layers and their subsequent

    reattachment. In addition, the windward pressures were significantly greater in the

    wall jet, due to the inverted velocity profile that decreased slightly with height.

    Further away from jet stagnation, X=D > 1:5; boundary layer development of thediverging wall jet flow occurs and consequently the pressures become more like those

    of conventional boundary layer tests. However, these regions experience much lowervelocities and consequently pressures and are thus not considered significant from a

    wind load design perspective.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732730

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    21/22

    This study has shown that mean pressure distributions on objects immersed in wall

    jets differ from those in conventional wind tunnel studies, however, as a realistic

    representation of wind pressures generated by thunderstorm downbursts, there

    remains considerable debate, not the least because the full-scale storms havesignificantly different kinematic structure compared to the quasi-steady wall jet

    simulation. Part B of this paper attempts to address these issues by describing results

    from the moving jet simulation, which successfully captured many of the

    characteristics of full-scale downbursts, including the transient gust front.

    Acknowledgements

    The authors would like to acknowledge support from the Wind Science and

    Engineering Research Center and a Seed Grant for Multidisciplinary Research from

    Texas Tech University.

    References

    [1] AS1170.2-1989, Australian standardminimum design loads on structuresPart 2: wind loads,

    Standards Australia, Homebush, 1989.

    [2] ASCE 7-98, Minimum design loads for buildings and other structures, Structural Engineering

    Institute of the American Society of Civil Engineers, Virginia, 1998.

    [3] ISO4354, Wind actions on structures, International Standards Organization, Geneva, Switzerland,1997.

    [4] J.D. Holmes, Modelling of extreme thunderstorm winds for wind loading of structures and risk

    assessment, Wind Engineering into the 21st Century: Proceedings of the 10th International

    Conference on Wind Engineering, Vol. 2. Balkema, Rotterdam, 1999, pp. 14091416.

    [5] C.W. Letchford, C. Mans, M.T. Chay, Thunderstormstheir importance in Wind Engineering, a

    case for the next generation wind tunnel, JAWE J. Wind Eng. 89 (2001) 3143.

    [6] T.T. Fujita, Andrews AFB Microburst, SMRP Research Paper 205, University of Chicago, 1983,

    38pp.

    [7] T.T. Fujita, Downburst: Microburst And Macroburst, University of Chicago Press, Chicago, IL,

    1985, 122pp.

    [8] M.R. Hjelmfelt, Structure and life cycle of microburst outflows observed in Colorado, J. Meteorol.

    27 (1988) 900927.

    [9] J.W. Wilson, R.D. Roberts, C. Kessinger, J. McCarty, Microburst wind structure and evaluation of

    Doppler radar for airport wind shear detection, J. Climate Appl. Meteorol. 23 (1984) 898915.

    [10] J.D. Holmes, Physical modelling of thunderstorm downdrafts by wind tunnel jet, Second AWES

    Workshop, Monash University, 2122 February 1992, pp. 2932.

    [11] G.S. Wood, K.C.S. Kwok, N.A. Motteram, D.F. Fletcher, Physical and numerical modeling of

    thunderstorm downbursts, Wind Engineering into the 21st Century: Proceedings of the 10th

    International Conference on Wind Engineering, Vol. 3, Balkema, Rotterdam, 1999, pp. 19191924.

    [12] C.W. Letchford, G. Illidge, Turbulence and topographic effects in simulated thunderstorm

    downdrafts by wind tunnel jet, Wind Engineering into the 21st Century: Proceedings of the 10th

    International Conference on Wind Engineering, Balkema, Rotterdam, Vol. 3, 1999, pp. 19071918.

    [13] W.D. Baines, Effects of velocity distribution on wind loads and flow patterns on a building,

    Proceedings, Symposium No. 16, Wind Effects on Buildings and Structures, England, 1963.

    [14] I.P. Castro, A.G. Robins, The flow around a surface-mounted cube in uniform and turbulent streams,

    J. Fluid Mech. 79 (2) (1977) 307335.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732 731

  • 7/29/2019 Pressure Distributions on a Cube in a Simultated Thunderstorm Downburts Part A

    22/22

    [15] A. Hunt, Wind-tunnel measurements of surface pressures on cubic building models at several scales,

    J. Wind Eng. Ind. Aerodyn. 10 (1982) 137163.

    [16] N. H.olscher, H.-J. Niemann, Towards quality assurance for wind tunnel tests: a comparative testing

    program of the Windtechnologische Gesellschaft, J. Wind Eng. Ind. Aerodyn. 7476 (1998) 599608.[17] S. Murakami, A. Mochida, 3-D numerical simulation of airflow around a cubic model by means of

    the ke model, J. Wind Eng. Ind. Aerodyn. 31 (1988) 283303.

    [18] D.A. Paterson, C.J. Apelt, Simulation of flow past a cube in a turbulent boundary layer, J. Wind Eng.

    Ind. Aerodyn. 35 (1990) 149176.

    [19] Y. Ogawa, S. Oikawa, K. Uehara, Field and wind tunnel studies of the flow and diffusion around a

    model cube1. Flow measurements, Atmos. Environ. 17 (6) (1983) 11451159.

    [20] P. Richards, R.P. Hoxey, L.J. Short, Wind pressures on a 6 m cube, Volume of Abstracts: Fourth

    International Colloquium on Bluff Body Aerodynamics and Applications, Ruhr University, Bochum,

    Germany, 2000, pp. 515518.

    [21] C.W. Letchford, C. Mans, Translational effects in modeling thunderstorm downbursts, Volume of

    Abstracts: Fourth International Colloquium on Bluff Body Aerodynamics and Applications, Ruhr

    University, Bochum, Germany, 2000, pp. 153156.[22] M.T. Chay, Simulation of thunderstorm downbursts, MSCE Thesis, Texas Tech University, 2001.

    [23] R.J. Cassar, Simulation of thunderstorm downdraft by a wind tunnel jet, Summer Vacation Project,

    CSIRO Australia, 1992, 44pp.

    [24] C.W. Letchford, M.T. Chay, Pressure distributions on a cube in a simulated thunderstorm

    downburstPart B: moving downburst observations, J. Wind Eng. Ind. Aerodyn. 90 (2002) 737.

    M.T. Chay, C.W. Letchford / J. Wind Eng. Ind. Aerodyn. 90 (2002) 711732732