22
Stack Gas Dispersion Measurements with Large Scale-PIV, Aspiration Probes and Light Scattering Techniques and Comparison with CFD une¸ sNakibo˘glu a,1 , Catherine Gorl´ e a,b ,Istv´anHorv´ath a , Jeroen van Beeck a , Bert Blocken c a Von Karman Institute for Fluid Dynamics, Waterloosesteenweg 72, 1640 Sint-Genesius-Rode, Belgium b University of Antwerp, Department of Physics, EMAT, Groenenborgerlaan 171, 2020 Antwerp, Belgium c Eindhoven University of Technology, Building Physics and Systems, P.O.Box 513, 5600MB Eindhoven, The Netherlands Abstract The main purpose of this research is to manage simultaneous measurement of ve- locity and concentration in large cross sections by recording and processing images of cloud structures to provide more detailed information for e.g. validation of CFD simulations. Dispersion from an isolated stack in an Atmospheric Boundary Layer (ABL) was chosen as the test case and investigated both experimentally and nu- merically in a wind tunnel. Large Scale-Particle Image Velocimetry (LS-PIV), which records cloud structures instead of individual particles, was used to obtain the veloc- ity field in a vertical plane. The concentration field was determined by two methods: Aspiration Probe (AP) measurements and Light Scattering Technique (LST). In the latter approach, the same set of images used in the LS-PIV was employed. The test case was also simulated using the CFD solver FLUENT 6.3. Comparison between AP measurements and CFD revealed that there is good agreement when using a turbulent Schmidt number of 0.4. For the LST measurements, a non-linear rela- tion between concentration and light intensity was observed and a hyperbolic-based function is proposed as correction function. After applying this correction function, a close agreement between CFD and LST measurements is obtained. Key words: Atmospheric Boundary Layer (ABL), Dispersion, Wind-tunnel fluid modeling, Particle Image Velocimetry (PIV), Light Scattering Technique (LST), Computational Fluid Dynamics (CFD), Turbulent Schmidt Number Accepted for publication in Atmospheric Environment 20 March 2009

Stack Gas Dispersion Measurements with Large Scale-PIV

  • Upload
    hatram

  • View
    229

  • Download
    1

Embed Size (px)

Citation preview

Page 1: Stack Gas Dispersion Measurements with Large Scale-PIV

Stack Gas Dispersion Measurements with

Large Scale-PIV, Aspiration Probes and Light

Scattering Techniques and Comparison with

CFD

Gunes Nakiboglu a,1, Catherine Gorle a,b, Istvan Horvath a,Jeroen van Beeck a, Bert Blocken c

aVon Karman Institute for Fluid Dynamics, Waterloosesteenweg 72, 1640

Sint-Genesius-Rode, Belgium

bUniversity of Antwerp, Department of Physics, EMAT, Groenenborgerlaan 171,

2020 Antwerp, Belgium

cEindhoven University of Technology, Building Physics and Systems, P.O.Box

513, 5600MB Eindhoven, The Netherlands

Abstract

The main purpose of this research is to manage simultaneous measurement of ve-locity and concentration in large cross sections by recording and processing imagesof cloud structures to provide more detailed information for e.g. validation of CFDsimulations. Dispersion from an isolated stack in an Atmospheric Boundary Layer(ABL) was chosen as the test case and investigated both experimentally and nu-merically in a wind tunnel. Large Scale-Particle Image Velocimetry (LS-PIV), whichrecords cloud structures instead of individual particles, was used to obtain the veloc-ity field in a vertical plane. The concentration field was determined by two methods:Aspiration Probe (AP) measurements and Light Scattering Technique (LST). In thelatter approach, the same set of images used in the LS-PIV was employed. The testcase was also simulated using the CFD solver FLUENT 6.3. Comparison betweenAP measurements and CFD revealed that there is good agreement when using aturbulent Schmidt number of 0.4. For the LST measurements, a non-linear rela-tion between concentration and light intensity was observed and a hyperbolic-basedfunction is proposed as correction function. After applying this correction function,a close agreement between CFD and LST measurements is obtained.

Key words: Atmospheric Boundary Layer (ABL), Dispersion, Wind-tunnel fluidmodeling, Particle Image Velocimetry (PIV), Light Scattering Technique (LST),Computational Fluid Dynamics (CFD), Turbulent Schmidt Number

Accepted for publication in Atmospheric Environment 20 March 2009

Page 2: Stack Gas Dispersion Measurements with Large Scale-PIV

1 Introduction

Research on air pollution problems has reached major interest in the lastfew decades due to increased environmental consciousness. As a result, thereis a growing demand for precise prediction of pollutant concentration, bothexperimentally and numerically. Direct field measurements and wind tunnelstudies are the two main branches of the experimental approach. The for-mer has the advantage of providing information on the real complexity of thephenomenon (Wilson and Lamb, 1994). The limitation however is that manygoverning parameters are simultaneously operative and uncontrollable (Gar-cia Sagrado et al., 2002). Wind tunnel studies, on the other hand, providea controllable physical environment where the effect of various parameterscan be isolated and investigated separately (Meroney et al., 1999), (Continiet al., 2006), (Gromke et al., 2008). The main disadvantage of wind tun-nel tests is that they are time consuming, costly and usually simplified interms of physical modelling. Analytical and numerical approaches can alsobe classified, based on the level of complexity. Semi-empirical, e.g Gaussianmodels (Turner, 1970), and operational models of integral nature, e.g. OSPM(Berkowicz, 2000) and ADMS-Urban (Carruthers et al., 1994), are simple andeasy-to-use tools, however they have a limited applicability. Simulations withComputational Fluid Dynamics (CFD) can offer solutions for geometricallycomplex situations. Although CFD is more demanding than the other calcu-lation approaches in terms of the required resources, it is in many cases stillless expensive than experimental approaches. A main advantage of CFD isthat a single calculation provides “whole-flow-field data”. In contrast to oper-ational models which have undergone many comprehensive formal evaluationsfor their fitness for purpose, CFD models do not have such an evaluationrecord in pollutant dispersion (Di Sabatino et al., 2007). There has currentlynot been sufficient validation to make a reasonable estimate of the accuracyof such concentration calculations (Riddle et al., 2004).

The primary goal of this paper is to manage simultaneous measurement ofvelocity and concentration at large scales where only cloud structures andnot individual particles are traced. In this context, measurements of stack gasdispersion were performed in a proper Atmospheric Boundary Layer (ABL)wind tunnel. The velocity field has been measured with Large Scale-Particle

Email addresses: [email protected] (Gunes Nakiboglu),[email protected] (Catherine Gorle), [email protected] (IstvanHorvath), [email protected] (Jeroen van Beeck), [email protected](Bert Blocken).1 Present address: Eindhoven University of Technology, Department of AppliedPhysics, Fluid Dynamics Laboratory, P.O.Box 513, 5600MB Eindhoven, TheNetherlands

2

Page 3: Stack Gas Dispersion Measurements with Large Scale-PIV

Image Velocimetry (LS-PIV). The concentration field has been obtained us-ing two different techniques, namely Aspiration Probe (AP) measurements andLight Scattering Technique (LST). The former is a well-established methodfor point concentration measurements (Maroteaux, 1991). The latter is lessestablished, however, it has the advantages of being a non-intrusive and a pla-nar measurement. For the LS-PIV and the LST, the same set of images wereemployed. Thus, in this sense, simultaneous measurement of velocity and con-centration was achieved. For a better understanding of the phenomenon andfor an extended comparison, the same case was also studied numerically. Inthe numerical part of the research the commercial CFD software FLUENT 6.3was used and approaches proposed in literature (Blocken et al., 2007), (Tom-inaga and Stathopoulos, 2007) were followed. Although studying dispersionof gases in a complex urban environment is essential and directly related tothe quality of life and safety of people living and working in such areas, it ismore appropriate to use a simple geometry for accuracy assessment purposesin initial studies (Riddle et al., 2004). Thus, in this research, gas dispersionfrom a single isolated stack in an ABL was studied, which is a well understoodcommon test case (Beychok, 1994), (Raza et al., 2001), (Blocken et al., 2008).

2 Wind Tunnel, Stack Model and Atmospheric Boundary Layer

Modeling

The experiments have been conducted in the Wind Gallery (Figure 1), anatmospheric boundary layer wind tunnel at the von Karman Institute (VKI).This facility is a low-speed, open-circuit, suction-type wind tunnel. It incor-porates an air inlet, fitted with honeycombs, meshes and a three-dimensionalcontraction. The test section is 1 m high, 1.3 m wide and 7 m long. A batteryof four ejectors is mounted at the back end to drive the airflow which can bevaried from 0.25 ms−1 to 1.2 ms−1. Four Irwin vortex generators (Irwin, 1981)were placed at the entrance of the development section. In addition to thespires, small cubic roughness elements with a specific size and arrangementwere fixed on the floor to control the properties of the induced boundary layer(Degroote, 2005).

The model of isolated stack used in the wind tunnel experiments has an inner(Di) and an outer (Do) diameter of 9 mm and 10 mm, respectively, and aheight (hs) of 100 mm. The velocity ratio (Vs/Us) at the stack is around 5.4.Us is the streamwise wind speed of the approach flow at stack height and Vs

is the stack exhaust velocity. For both AP and LS-PIV/LST measurements,the value of Vs leads to stack Reynolds numbers (Res) which are above the

3

Page 4: Stack Gas Dispersion Measurements with Large Scale-PIV

Fig. 1. Wind Gallery wind tunnel at von Karman Institute

Res limit of 2000 proposed by Snyder (1981), which is necessary to ensureturbulent stack flow.

Measurements of the vertical approach flow profiles were performed at 0.1m upstream of the stack with a boundary layer hot-wire probe which was cal-ibrated with Laser-Doppler-Velocimetry (LDV) for very low velocities downto 0.05 ms−1. For data acquisition the KUSB-3102 module of Keithley Instru-ments, Inc. was used together with TestPointTM software. The profiles wereobtained from approximately 60 measurement locations along a vertical line,with a high resolution near the wall region. A sampling rate of 20Hz and sam-pling time of 60s were used. This rather long sampling time was chosen toeliminate time localized fluctuations of the wind tunnel.

Initial attention was given to proper modeling of the turbulent boundary layer,which requires validation of a minimum roughness Reynolds number (Re∗).

Re∗ =u∗z0

ν=

0.057ms−1× 0.00144m

1.5 × 10−5m2s−1= 5.47 (1)

where the friction velocity, u∗, was determined from the logarithmic verticalprofile of the mean wind speed in the wind tunnel. The aerodynamic rough-ness length, z0, was determined using the height of the real roughness elements(Ks = 18mm) and the linear relation proposed by Rau et al. (1991) for ho-mogeneous terrains:

z0 = 0.08Ks (2)

Note that Ks is not the equivalent sand-grain roughness height (ks) but theactual height of the individual roughness elements in the wind tunnel. Thefull-scale aerodynamic roughness length (z0) is about 0.3 m. According toPlate (1982), the roughness Reynolds number should be higher than 5 to en-sure aerodynamically rough flow, but Snyder and Castro (1998) later showed

4

Page 5: Stack Gas Dispersion Measurements with Large Scale-PIV

that Re∗ = 1 is large enough. The current value of Re∗ = 5.47 satisfies thesecriteria. Thus, at the end of development section, an aerodynamically roughflow is obtained. Measured mean wind velocity (U/Us) and longitudinal tur-bulence intensity (σu/U) profiles are presented in Figure 2. These profiles are

σu / Uz

[m]

0 0.1 0.2 0.3 0.4 0.5 0.60

0.1

0.2

0.3

0.4

0.5

U / Us

z[m

]

0.2 0.4 0.6 0.8 1 1.2 1.40

0.1

0.2

0.3

0.4

0.5

Fig. 2. Mean velocity and longitudinal turbulence intensity profiles

representative of an ABL and will later be employed as boundary conditionsin the numerical modeling of the test case.

The typical turbulent length scale in the wind tunnel was also calculated.The Lagrangian velocity autocorrelation function (Pope, 2000) was computedusing a time series that was obtained from a single spatial position in thefree stream. Then Taylor’s frozen turbulence approximation (Lumley, 1964)was used to relate space and time. That is, a time lag and a mean-convectingvelocity give a spatial separation which is around 0.39m. These large scalescorrespond to the typical length scale of the Irwin vortex generators at theentrance of the development section.

3 Measurement Techniques

The following sections describe the different measurement techniques used, i.e.Large Scale-PIV for the velocity measurements and Aspiration Probe measure-ments and Light Scattering Technique for the concentration measurements.

3.1 Velocity Measurements: Large Scale-PIV (LS-PIV)

To determine the velocity field downstream of the stack, LS-PIV has been used,where cloud structures are visualized instead of individual particles. The ex-perimental set-up is presented in Figure 3 (a). The flow was illuminated with

5

Page 6: Stack Gas Dispersion Measurements with Large Scale-PIV

a double cavity pulsed Mini:YAG laser (New Wave Research) which has apulse power of 200mJ. One spherical and one cylindrical lens with properfocal lengths were used to achieve a focalized light sheet. During the exper-iments a PCO Sensicam-12 bit gray scale camera (PCO Imaging) was usedwith a resolution of 1280 × 832. The synchronization was managed using theMotionProX timing hub signal generator (IDT Inc).

Throughout the experimental campaign, depending on the case, either oneor two VKI built-in smoke generators have been used. One of these smokegenerators was employed to seed the upstream flow and the other was usedto create a plume from the stack. The working principle of these devices isas follows (Scarano and Riethmuller, 1998): a nozzle sprays oil droplets on aheated plate at 180◦C by means of compressed air at 1.8 bar. The oil evap-orated from the plate exits the chamber and enters the wind tunnel throughflexible pipes. The created particles have a typical diameter in the range 1-3µm. The outflow from the smoke generator that is used to create the plumeis first mixed with air before entering the wind tunnel through the stack, asshown in Figure 3. The field of view was approximately 0.3 m × 0.4 m which

Fig. 3. Experimental set-ups for Large Scale-PIV:(a) and Aspiration Probe:(b) mea-surements

is quite large compared to standard PIV applications. A typical instantaneousimage, acquired during the experiments, is presented in Figure 4. For the dig-ital image processing of the PIV images, the WIDIM (Window DisplacementIterative Multigrid) algorithm developed by Scarano (1997) was used. Theaverage and RMS flow fields were obtained using 500 image couples per case.To improve the process of averaging, a technique called signal-to-noise ratiofiltered averaging, was used (Nakiboglu, 2008). In this approach, only vali-dated vectors from each instantaneous field were used in the averaging. Thethreshold for signal-to-noise ratio for validation was chosen as 1.5, which isthe default value of WIDIM.

6

Page 7: Stack Gas Dispersion Measurements with Large Scale-PIV

Fig. 4. A typical instantaneous image, acquired during the experiments(Do = 10mm, Vs/Us = 5.4)

3.2 Concentration Measurements: Aspiration Probe Measurements

Measurements of the instantaneous concentration were first performed withan aspiration probe (also known as hot-wire-sonic nozzle probe) manufacturedat the VKI. This method is a reliable intrusive point measurement technique(Maroteaux, 1991). The working principle is based on the properties of a con-stant temperature hot-wire sensor combined with a sonic nozzle. The hot-wireis sensitive to the velocity, temperature, pressure and physical properties of thefluid flowing around it. The sonic nozzle makes the hot-wire output insensitiveto the free stream velocity. In addition, in the present experiments, the changein temperature and pressure is negligible. Consequently, the hot-wire voltagedepends only on the physical properties of the fluid, meaning its composition.Therefore the device is capable of measuring tracer gas concentrations.

Due to its high traceability in air, Helium was used as tracer gas. By using acalibration box, the gas mixture ejected from the stack had a known sourceconcentration (Cs). Throughout the experiments a mixture of air (90%) andHelium (10% by mass) was introduced from the stack, which exhibits approx-imately neutrally buoyant plume characteristics. A sampling needle, namelyPrandtl tube of 1mm diameter was used together with a vacuum pump tochoke the flow in the sonic hole as shown in Figure 3 (b). Helium gas measure-ments were performed at 38 locations along seven vertical lines downstreamof the stack. For each point, measurements were taken during 60s with a sam-pling frequency of 50Hz.

7

Page 8: Stack Gas Dispersion Measurements with Large Scale-PIV

3.3 Concentration Measurements: Light Scattering Technique

More complete data sets for pollutant dispersion can be obtained by measur-ing concentrations in a plane. 2D laser optical concentration measurementsare very favorable in this sense (Garcia Sagrado et al., 2002). In general, theintensity of the light scattered from the particles is proportional to the lo-cal particle density in the light sheet (van de Hulst, 1957). When introducingseeded air from the stack into the unseeded air in the wind tunnel, the numberof scattering particles in each unit volume of the flow field will be reduced asthe control volume moves downstream of the seeding source and the intensityof the scattered light will diminish proportionally. In the present study, thesame set of images that was used for the PIV study was also employed for theLST measurements. Thus, the velocity and the average concentration fieldwere measured simultaneously.

The images were processed on averaging windows of 10 × 10 pixel size us-ing the following procedure:

(1) The relative concentration was calculated by eliminating the effect ofbackground light and non-uniformity of the illumination using the fol-lowing relation:

C

Cs

=I − Ib

Is − (Ib)s

(I0)s

I0

(3)

Where C, Cs and I, Is are the concentration and light intensity at a givenlocation and at the exit of the stack respectively. Ib and (Ib)s refer to thebackground light intensity. I0 and (I0)s refer to the light intensity of auniformly seeded background.

(2) When there is an excessive amount of particles in the flow, multiple scat-tering occurs, leading to an attenuation of the scattered light intensityreaching the camera. Equation 3 needs to be corrected through a non-linear function, affecting mostly the higher concentration regions. Fur-thermore, this function should preserve the extremities of the relativeconcentration at 0 and 1 and the function should be concave in nature.In addition the function should have a finite gradient at 0. Hyperbolicfunctions in general satisfy these requirements and the following equationwas derived for this purpose:

CReal

(CReal)s

= 1 −tanh

(

β(

1 −CLST

(CLST )s

))

tanh(β)(4)

where CReal, (CReal)s and CLST , (CLST )s are the real concentration and

8

Page 9: Stack Gas Dispersion Measurements with Large Scale-PIV

the concentration estimated from the LST measurements at a chosenlocation and at the exit of the stack, respectively. The coefficient β hasto be determined from calibration, which can be performed based on alimited number of AP point measurements or based on validated CFDsimulations.

4 Numerical simulations

CFD models are increasingly used to investigate atmospheric dispersion pro-cesses in the lower region of the ABL (Chang and Meroney, 2003). In thepresent study the CFD simulations were used to determine the coefficient βin equation 4. For this purpose, a first set of simulations was performed tovalidate the CFD simulations with the AP measurements. Afterwards thesevalidated simulation settings were used for performing simulations that couldbe compared with the LST measurements. An accurate prediction of the dis-persion phenomenon highly depends on successful ABL modeling (Blocken etal., 2007). Thus, initial attention in the numerical part of the research wasgiven to the simulation of the ABL, as created in the wind tunnel. Secondly,the dispersion from the single isolated stack was investigated using the veri-fied ABL and the simulations were compared with the LS-PIV, AP and LSTmeasurements.

4.1 Simulation of the atmospheric boundary layer

The common practice in most CFD investigations is that the upstream terrainof the region of interest is not modeled, but instead fully developed inlet pro-files of velocity and turbulent quantities are specified at the inlet plane of thecomputational domain. These profiles should be representative of the upstreamregion that is not included in the computational domain. A problematic partis the definition of proper wall functions at the bottom of the domain suchthat the vertical profiles of the mean wind speed and turbulence as defined atthe inlet have a zero streamwise gradient. This is referred to as a horizontallyhomogeneous ABL. The difficulty of achieving such a boundary layer has al-ready been highlighted in the literature (Riddle et al., 2004), (Blocken andCarmeliet, 2006). Thus, first a set of 2D simulations was performed to achievea horizontally homogeneous ABL using the method proposed by Blocken etal. (2007).

In this study the standard k-ǫ turbulence model (Jones and Launder, 1972)was used. As the turbulence model assumes isotropic turbulence, the reportedturbulence intensity profile in Figure 2 was converted to turbulent kinetic

9

Page 10: Stack Gas Dispersion Measurements with Large Scale-PIV

energy (k) as follows:

k(z) =3

2σu

2 (5)

where σu is the standard deviation of the velocity fluctuations in the x-direction. The inlet profile for the turbulence dissipation rate ǫ was determinedusing the following formula (Tominaga et al., 2008), (Gorle et al., 2009):

ǫ =√

Cµk(z)dU

dz(6)

which implies the equilibrium of turbulence production and dissipation. Thedefault value of 0.09 was used for Cµ. For the wall boundary condition therelation proposed by Blocken et al. (2007) which brings the wall functions(Launder and Spalding, 1974) in equilibrium with the velocity inlet profilewas employed. For FLUENT 6.3 this relation is given as

ks,ABL =9.793z0

CKs

(7)

where ks,ABL is the equivalent sand-grain roughness height as defined in FLU-ENT for the rough wall function and CKs is a constant, required for the wallfunction. There are two constraints which should be taken into account in thisapproach as highlighted in the FLUENT manual. First, ks,ABL should be lessthan the half of the height of the wall-adjacent cell. Second, CKs should beless than one. However, if both constraints are respected even for the largestpossible CKs value, it is not possible to satisfy Equation 7. Therefore the re-quirement of limiting CKs to 1 was ignored and higher values were definedthrough a user-defined function (UDF).

In these 2D simulations for testing the horizontal homogeneity of the ABL,an empty domain of length 0.7m and height of 0.6m was used. To generatethe grid GAMBIT 2.3.16. was employed. A structured grid was used whichwas highly clustered close to ground level, to have enough resolution in thelower range of ABL. The height of the wall-adjacent cell was set at 0.008m.With z0 = 0.00144 m and ks = 0.004 m, CKs was set as 3.53 through Equation7. A total of 6600 quadrilateral cells (88 × 75) were employed. The SIMPLEalgorithm was used in the simulations and all discretization schemes were setto second order (Franke et al., 2004). The iterations were terminated when allresiduals had dropped at least 8 orders of magnitude. The resulting velocityand turbulence properties at the inlet and the outlet of the computational do-main are presented in Figure 5. It can be noticed that horizontal homogeneitywas achieved for the velocity profile but not for the turbulence properties. The

10

Page 11: Stack Gas Dispersion Measurements with Large Scale-PIV

µt [kg/m-s]

z[m

]

0.005 0.01 0.015 0.020

0.1

0.2

0.3

0.4

Turbulent Viscosity Profile

ε [m 2/s3]

z[m

]

0 0.05 0.1 0.15 0.20

0.1

0.2

0.3

0.4

Turbulent Dissipation Profile

u [m/s]

z[m

]

0 0.2 0.4 0.6 0.8 10

0.1

0.2

0.3

0.4Experimental data (CFD Inlet)CFD outlet

Velocity Profile

k [m 2/s2]

z[m

]

0 0.01 0.02 0.030

0.1

0.2

0.3

0.4

Turbulent Kinetic Energy Profile

Fig. 5. Comparison of velocity and turbulence properties between the inlet (exper-imental data) and outlet planes

horizontal homogeneity of the turbulence properties could be improved by ad-justing the inlet profiles and the constants Cµ and σǫ in the k-ǫ turbulencemodel as described by Gorle et al. (2009). In the 3D simulations of the stackgas dispersion however, the turbulent kinetic energy is only used indirectlythrough the turbulent viscosity, µt = ρCµk

2/ǫ, and the horizontal variationhereof is limited as shown in Figure 5.

4.2 Simulation of stack gas dispersion

4.2.1 Computational model and boundary conditions

The 3D domain used for the simulations had a length of 0.7m, a width of0.65m and a height of 0.6m. The stack was located at 0.1m from the inlet andhad a height of 0.1m. The model had a symmetry plane through the centerlineof the stack in the streamwise direction to decrease the computational load.The grid was constructed by gridding the ground surface with unstructuredquad cells, which were clustered close to the stack by means of size functions.Using the Cooper scheme (Fluent Inc., 2006) this face grid was extruded along

11

Page 12: Stack Gas Dispersion Measurements with Large Scale-PIV

the height of the domain to obtain 3D cells. The height of the wall-adjacentcells was identical to the one used for the 2D ABL simulations. Approximately350000 cells were created with a skewness angle less than 0.4 for all cells. Thecomputational domain and the grid within are presented in Figure 6. In thesimulations which are compared with the AP measurements, the stack exitvelocity was determined from the measured value and that for the comparisonwith the LST was determined based on fitting the plume centerline deflection.Simulation parameters that are not explicitly mentioned here were taken thesame as for the 2D test simulations.

0

Zoom View

x [m]

00.2

0.40.6

y [m]

0

0.2

0.4

0.6

z[m

]

0

0.2

0.4

0.6

3D View

Fig. 6. Grid that is created in GAMBIT 2.3.16. for 3D stack gas dispersion

4.2.2 Dispersion Modelling

To determine the concentration distribution, species transport calculationswere performed by solving the scalar advection-diffusion equation, employingFick’s law of diffusion. When the flow is turbulent, Fick’s law of diffusion takesthe following form;

Ji = −(ρDi,m + ρDt)∇Yi (8)

where Ji is the diffusion flux of chemical species, i; in this case Helium, Yi

is the mass fraction of the species, Di,m and Dt are the molecular mass dif-fusion coefficient and the turbulent mass diffusion coefficient, respectively. Inturbulent flows the dominant term for diffusion is the turbulent mass diffusioncoefficient which is expressed as follows,

Dt =µt

ρSct

(9)

12

Page 13: Stack Gas Dispersion Measurements with Large Scale-PIV

where µt is the eddy viscosity and Sct is the turbulent Schmidt number. Thereported optimum values for Sct in the literature are widely distributed in therange of 0.2-1.3 and the specific value selected has a significant effect on theresults (Tominaga and Stathopoulos, 2007). Early studies of CFD modelingof turbulent mass transport (Spalding, 1971), (Launder, 1978) pointed outthat values of Sct between 0.7 and 0.9 provided the best agreement withexperimental results. However, recent studies comparing CFD with provedintegral models and experiments for dispersion in the ABL (Riddle et al.,2004), (Di Sabatino et al., 2007), (Blocken et al., 2008) showed that increasingthe plume dispersion by reducing the Sct to values around 0.3 and 0.4 improvesthe predicted concentration levels. Thus, in this study, different Sct numbersas proposed in the literature are investigated.

5 Comparison of results

A comparison of the results obtained from the different measurements andthe CFD simulations is presented for the regions where both experimentaland numerical data are available. This region starts at the stack and extendsto 30 stack diameters (SD) downstream. First, the velocity fields obtainedfrom LS-PIV and CFD are addressed. Second, a comparison between the APconcentration measurements and the CFD results is presented and finally thecomparison between the LST measurements and CFD is discussed.

5.1 Velocity field

First, a comparison between different seeding configurations as mentioned insection 3.1 was performed. The results showed that seeding the flow not onlyfrom the stack but also from the free stream improves the valid vector ratio(VVR), as expected. The flow seeded from the stack only had a VVR of 73.6%.This rather low VVR is mainly due to the region upstream of the stack wherethere is no seeding and the area very close to the exit of the stack where there isexcessive seeding. Introducing seeding in the free stream improved the resultsby as much as 13%. However, upstream seeding is not an option for opticalconcentration measurement techniques because it would be a second sourceof concentration. Thus, it is important to know how successfully the velocityfield can be obtained by seeding the flow only from the stack. In Figure 7the velocity profiles that were obtained by means of PIV measurements withseeding from the stack only and seeding from the stack and the upstream arecompared with each other. It is observed that the difference in the velocitiesbetween the two seeding configurations decreases when moving downstream.Behind 7.5 SD downstream the maximum and average differences are already

13

Page 14: Stack Gas Dispersion Measurements with Large Scale-PIV

U [m/s]

z[m

]

0 0.5 1 1.5 2

0.1

0.15

0.2

0.25

Stack OnlyUpstream&Stack

x = 0 SD

U [m/s]

z[m

]

0 0.5 1 1.5 2

0.1

0.15

0.2

0.25

x = SD

U [m/s]z

[m]

0 0.5 1 1.5 2

0.1

0.15

0.2

0.25

x = 2.5 SD

U [m/s]

z[m

]

0 0.5 1 1.5 2

0.1

0.15

0.2

0.25

x = 7.5 SD

U [m/s]

z[m

]

0 0.5 1 1.5 2

0.1

0.15

0.2

0.25

x = 5 SD

Fig. 7. Comparison of mean streamwise velocity profiles at various distances down-stream of the stack with different seeding configurations

less than 8% and 2%, respectively. Thus, seeding the flow only from the stackis sufficient to get the velocity information and simultaneously allows to ob-tain the concentration field by means of LST.

It should also be highlighted that from the PIV images it was observed thatdense clouds of particles coming out of the stack move like blocks. As a result,the motion of cloud structures can not be captured which leads to very low(< 0.5) signal-to-noise ratios in that region. Thus, both seeding configurationsare not reliable in resolving the velocity at the stack exit and in the regionclose to the stack.

Figures 8 and 9 show that there is a good agreement between CFD and PIVresults, not only in the mean but also in fluctuating component of the velocityfield, especially after 2.5 SD downstream of the stack. To be able to comparethe fluctuating components of velocity that were obtained during the PIVmeasurements in streamwise (σu) and lateral directions (σv) with isotropicturbulent CFD simulations; the isotropic velocity fluctuation (σiso) was cal-culated through the following relation: σiso = [(σu + σv)/2]0.5. It is seen fromFigure 9 that the difference between the horizontal and vertical fluctuationsof the velocity is quite large at the cross section x = 2.5 SD. Thus, the as-sumption of isotropic turbulence in the CFD simulations with the standardk-ǫ model is less appropriate at this position. Together with the low signal-to-noise ratios in the proximity of the stack, this can be the main reason of thediscrepancy between the simulations and experiments in the region close tostack. The region close to stack could be better resolved in the experiments,simply by decreasing the density of the particles coming out of the stack. Thiswould, however, decrease the amount of scattered light, which is unfavourablefor the LST measurements.

14

Page 15: Stack Gas Dispersion Measurements with Large Scale-PIV

U [m/s]

z[m

]

0 0.5 1 1.5 20.1

0.15

0.2

0.25

0.3PIVCFD

x = 2.5 SD

U[m/s]0 0.5 1 1.5 2

0.1

0.15

0.2

0.25

0.3

x =7.5 SD

U [m/s]0 0.5 1 1.5 2

0.1

0.15

0.2

0.25

0.3

x = 12.5 SD

U [m/s]0 0.5 1 1.5 2

0.1

0.15

0.2

0.25

0.3

x =17.5 SD

U [m/s]0 0.5 1 1.5 2

0.1

0.15

0.2

0.25

0.3

x = 22.5 SD

Fig. 8. Comparison of mean streamwise velocity profiles at various distances down-stream of the stack obtained by CFD and PIV

σ [m/s]

z[m

]

0 0.2 0.4 0.60.1

0.15

0.2

0.25

0.3CFDPIV - σu

PIV - σv

PIV - σiso

x = 2.5 SD

σ [m/s]

0 0.2 0.4 0.60.1

0.15

0.2

0.25

0.3x = 7.5 SD

σ [m/s]

0 0.2 0.4 0.60.1

0.15

0.2

0.25

0.3x = 12.5 SD

σ [m/s]

0 0.2 0.4 0.60.1

0.15

0.2

0.25

0.3x = 17.5 SD

σ [m/s]

0 0.2 0.4 0.60.1

0.15

0.2

0.25

0.3x = 22.5 SD

Fig. 9. Comparison of σ profiles at various distances downstream of the stack ob-tained by CFD and PIV

5.2 Concentration field from AP measurements and CFD

The concentration fields obtained by the AP measurements and the CFDsimulations are presented in Figure 10, for three different turbulent Schmidtnumbers: 0.4, 0.7 and 1. It is noticed that the lower turbulent Schmidt num-ber 0.4 provides a better agreement with the experimental data. This is in

15

Page 16: Stack Gas Dispersion Measurements with Large Scale-PIV

agreement with the findings of Di Sabatino et al. (2007) and Blocken et al.(2008). A point-wise comparison of the experimental and CFD data is pre-

C/Cs

z[m

]

0 0.4 0.80

0.1

0.2

0.3

Turb. Schmidt 0.4Turb. Schmidt 0.7Turb. Schmidt 1Aspiration Probe

x = 0 mm (Stack Plane)

C/Cs

z[m

]

0 0.40

0.1

0.2

0.3

x = 25mm (2.5 SD)

C/Cs

z[m

]0 0.4

0

0.1

0.2

0.3

x = 50mm (5 SD)

C/Cs

z[m

]

0 0.40

0.1

0.2

0.3

x = 75mm (7.5 SD)

C/Cs

z[m

]

0 0.40

0.1

0.2

0.3

x = 100mm (10 SD)

Fig. 10. Comparison of relative Helium concentrations at various distances from thestack determined by AP measurements and CFD (with different Sct numbers)

sented in Figure 11. The concentrations measured with the AP and obtainedfrom the CFD simulations are in agreement within a relative concentration(C/Cs) band of 0.04, which is considered a very close agreement.

C/Cs (Experiment)

C/C

s(C

FD

)

0 0.05 0.1 0.15 0.20

0.05

0.1

0.15

0.2

Sct = 0.4Sct = 0.7Sct = 1Exp = CFD

Fig. 11. Point-wise comparison of experimental data obtained by AP measurementsand CFD (with different Sct numbers)

5.3 Concentration field from LST measurements and CFD

A comparison of the concentration field measured with LST and calculatedwith CFD is presented in Figure 12. The figure includes “LST Raw”, which

16

Page 17: Stack Gas Dispersion Measurements with Large Scale-PIV

is the relative concentration corrected for the background light and non-homogeneity of the illumination using Equation 3 and “LST Tanh”, whichincludes the additional correction for the failure of independent light scatter-ing according to Equation 4. The coefficient β in this equation was determinedas 1.7 for this set of experiments based on the CFD simulations validated withthe AP measurements.

C/Cs

0 0.2 0.4

0.1

0.15

0.2

0.25

0.3

x = 17.5 SD

C/Cs

0 0.2 0.4

0.1

0.15

0.2

0.25

0.3

x = 22.5 SD

C/Cs

0 0.2 0.4

0.1

0.15

0.2

0.25

0.3

x = 10 SD

C/Cs

z[m

]

0 0.4

0.1

0.15

0.2

0.25

0.3

CFDLST RawLST Tanh

x = 1 SD

C/Cs

0 0.2 0.4

0.1

0.15

0.2

0.25

0.3

x = 5 SD

Fig. 12. Comparison of concentration fields obtained by LST measurements andCFD. “LST Raw” and “LST Tanh” stand for the raw data and the data correctedwith Equation 4, respectively.

It is noted that “LST Raw” overestimates the concentration. This is explainedby the fact that, at the stack exit, an excessive amount of particles is present,which does not result in independent scattering and thereby leads to an un-derestimated concentration at this location. Since the stack exit was chosen asthe reference location for the determination of the relative concentration ac-cording to Equation 3, the relative concentration will be overestimated. Whenintroducing the correction as proposed by Equation 4, this overestimation iseliminated. It is shown that the proposed correction is successful both at highand at low concentration levels.

6 Discussion

LS-PIV/LST was successfully implemented to obtain simultaneously the ve-locity and concentration fields. However, a few limitations need to be men-tioned. In the LS-PIV/LST measurements, the direct determination of thestack exit velocity was not possible. The rotameter could not be used becausethe smoke in the flow would lead to inaccurate flowrate measurements. Thedetermination of the exit velocity by LS-PIV was also not possible, due to thedense seeding coming out of the stack; the block motion of the plume hinders

17

Page 18: Stack Gas Dispersion Measurements with Large Scale-PIV

the formation of cloud structures. It is important that future research focuseson accurate source identification in case of seeding of the stack by smoke.Non-intrusive techniques like conventional PIV or LDV could be envisaged.

It should be noted that, because of the inability to correctly determine thestack exit velocity during the LS-PIV experiments, somewhat different mo-mentum ratios were used for AP and LS-PIV/LST measurements. Conse-quently, a direct comparison between both sets of results was not possible.For this reason, the AP measurements were used to validate the first set ofCFD simulations and to determine the optimal turbulent Schmidt number.The second set of CFD simulations, for the LS-PIV/LST case, were performedusing this same turbulent Schmidt number.

The buoyancy effect due to the higher temperature of the smoke could poten-tially introduce another problem. In the present study this effect was limitedbecause the outflow from the smoke generator used to create the plume wasfirst mixed with air before entering the wind tunnel through the stack. Thisdecreased the temperature of the mixture to achieve a nearly naturally buoy-ant plume. A preliminary test was conducted with a horizontal jet to observethe buoyancy characteristic of the plume. The jet showed no vertical deviationwithin a downstream distance of 1m, indicating that the smoke was probablycompensating for the buoyancy due to temperature. At a further distance, thesmoke rained out, causing the jet to go down.

Since the study was performed at model scale, the stack Reynolds number, aswell as the Reynolds number for the flow around the stack, are much smallerthan they would be in full-scale conditions. Since the main purpose of thisstudy was the demonstration of simultaneous velocity and concentration mea-surements and CFD validation based on these measurements, this is not con-sidered problematic for this study.

7 Conclusion

Gaseous pollutant dispersion from a single, isolated stack in an AtmosphericBoundary Layer (ABL) has been investigated by means of experiments in theWind Gallery facility at the VKI and numerical simulations with the CFDcode FLUENT 6.3. In the experimental part the velocity field was measuredwith Large Scale-Particle Image Velocimetry (LS-PIV) where the movementof structures was correlated instead of particles. The concentration field wasobtained by aspiration probe (AP) measurements and the Light Scattering

18

Page 19: Stack Gas Dispersion Measurements with Large Scale-PIV

Technique (LST). In the numerical part, initial attention was given to propermodeling of the atmospheric boundary layer, to allow correct modeling of thegas dispersion in a second phase.

LS-PIV was successfully implemented to obtain the velocity field. Differentseeding configurations were studied and it was concluded that introducing par-ticles only from the stack is sufficient to obtain the velocity field. This seedingconfiguration is necessary to allow for simultaneous concentration measure-ments by LST. It was also noticed that the determination of the velocity isproblematic in the region close to the stack. The dense seeding coming out ofthe stack leads to a block motion of the plume which hinders the formationof cloud structures. Thus, it was not possible to obtain reliable informationabout the velocity field in the region between the stack and 2.5 stack diame-ters downstream.

Concentration measurements with AP confirmed the reliability of this method.However, it has the inherent limitation of being a point measurement tech-nique. It was shown that the same set of images used for LS-PIV can also beused for concentration measurements through LST which will lead to 2D fieldinformation instead of a point measurement. A non-linear relation between thescattered light and concentration was observed. A hyperbolic-based functionwas proposed for the correction.

The velocity profile of the ABL that was generated in the wind tunnel wassuccessfully modeled in the CFD simulations using the proposed approaches inthe literature. The importance of the turbulent Schmidt number for dispersionsimulations was confirmed. Agreeing with other recent studies, it was shownthat decreasing the Sct to 0.4 improves the prediction of the concentrationsand yields a very close agreement of the CFD results with both AP and LSTmeasurements.

Acknowledgments

The authors wish to thank B. Toth, F. Tomasoni, R. Theunissen and J. Mis-sart for their help in setting up the experiments and post-processing.

References

Berkowicz, R., 2000. OSPM-a parameterised street pollution model. Journalof Environmental Monitoring and Assessment, 65, 323-331.

19

Page 20: Stack Gas Dispersion Measurements with Large Scale-PIV

Beychok, M. R., 1994. Fundamentals of stack gas dispersion. 3rd edition,Irvine, California.

Blocken, B., Carmeliet, J., 2006. The influence of the wind-blocking effect bya building on its wind-driven rain exposure. Journal of Wind Engineeringand Industrial Aerodynamics, 94(2), 101-127.

Blocken, B., Stathopoulos T., Carmeliet, J., 2007. CFD simulation of theatmospheric boundary layer: wall function problems. Atmospheric Environ-ment, 41, 238-252.

Blocken, B., Stathopoulos T., Saathoff, P., Wang, X., 2008. Numerical evalu-ation of pollutant dispersion in the built environment: Comparison betweenmodels and experiments. Journal of Wind Engineering and Industrial Aero-dynamics, 96(10-11), 1817-1831.

Carruthers, D.J., Holroyd, R.J., Hunt, J.C.R., Weng, W.S., Robins, A.G., Ap-sley, D.D., Thompson, D.J., Smith, F.B., 1994. UK-ADMS: A new approachto modeling dispersion in the earth’s atmospheric boundary layer. Journalof Wind Engineering and Industrial Aerodynamics, 52, 139-153.

Chang, C., Meroney, R., 2003. Concentration and flow distributions in ur-ban street canyons: wind tunnel and computational data. Journal of WindEngineering and Industrial Aerodynamics, 91, 1141-1154.

Contini, D., Hayden, P., Robins, A., 2006. Concentration field and turbulentfluxes during the mixing of two buoyant plumes. Atmospheric Environment,40(40), 7842-7857.

Degroote, Q., 2005. Project Report (UCL/VKI): Etude experimentale deseffects de la densite sur la dispersion de polluants. Sint-Genesius-Rode: vonKarman Institute for Fluid Dynamics.

Adrian, R. J., Durao, D. F. G., Durst, F., Heitor, M. V., Maeda, M., Whitelaw,J. H., 1998, Laser Techniques Applied to Fluid Mechanics: Selected Papersfrom the 9th International Symposium, Lisbon, Portugal, Sprinker.

Garcia Sagrado A., van Beeck, J., Rambaud, P., Olivari, D., 2002. Numeri-cal and experimental modeling of pollutant dispersion in a street canyon.Journal of Wind Engineering and Industrial Aerodynamics, 90, 321-339.

Gorle, C., van Beeck, J., Rambaud, P., Van Tendeloo, G., 2009. CFD modelingof small particle dispersion: the influence of the turbulence kinetic energyin the atmospheric boundary layer, Atmospheric Environment, 43, 673-681.

Gromke, C., Buccolieri, R., Di Sabatino, S., Ruck, B., 2008. Dispersion studyin a street canyon with tree planting by means of wind tunnel and nu-merical investigations - Evaluation of CFD data with experimental data,Atmospheric Environment, 42, 8640-8650.

Fluent 6.3 User’s Guide, 2006. FLUENT INC.Franke, J., Hirsch, C., Jensen, A.G., Krus, H.W., Schatzmann, M., Westbury,

P.S., Miles, S.D., Wisse, J.A. and Wright, N.G., 2004. Recommendationson the use of CFD in wind Engineering. Proceedings of the Internationalconference on Urban Wind Engineering and Building Aerodynamics, in: vanBeeck J.(Ed.) COST Action C14, Impact of Wind and Storm on City LifeBuilt Environment, von Karman Institute, Sint-Genesius-Rode, Belgium,

20

Page 21: Stack Gas Dispersion Measurements with Large Scale-PIV

5-7 May 2004.Irwin, H. A., 1981. The design of spires for wind simulation. Journal of Wind

Engineering and Industrial Aerodynamics, 7, 361-366.Jones, W. P., Launder, B.E., 1972. The prediction of laminarization with a

two-equation model of turbulence. International Journal of Heat and MassTransfer, 15, 2, 301-314.

Launder, B.E., Spalding, D.B., 1974. The numerical computation of turbulentflows. Computer Methods in Applied Mechanics and Engineering, 3, 2, 269-289.

Launder, B.E., 1978. Heat and Mass Transport. In: Bradshaw, P. (Ed.), Topicsin Applied Physics, Turbulence, vol. 12, Springer, Berlin.

Lumley, J. L., 1964. Interpretation of time spectra measured in high-intensityshear flows. The Physics of Fluids, 8(6), 1056-1062.

Maroteaux, D., Maroteaux, F., Murat, M., 1991. Experimental study of ahot wire, sonic nozzle probe for concentration measurements. Review ofScientific Instruments, 62,1057-1062.

Meroney, R. N., Leitl, B. M., Rafailidis, S., Schatzmann, M., 1999. Windtunnel and numerical modeling of flow and dispersion about several buildingshapes. Journal of Wind Engineering and Industrial Aerodynamics, 81(1-3),333-345.

Nakiboglu, G., 2008. Project Report: Velocity and concentration measure-ments of a stack gas dispersion in an atmospheric boundary layer. Sint-Genesius-Rode: von Karman Institute for Fluid Dynamics.

Plate, E., 1982. Engineering Meteorology, Elsevier Sci. Pub. Co., Amsterdam,The Netherlands.

Pope, S. B., 2000. Turbulent Flows. Cambridge University Press.Rau, M., Wang, Z. Y., Plate, E. J., Keiser, R., 1996. Scale effect in wind tunnel

modeling. Journal of Wind Engineering, 61,113-130.Raza, S. S., Avila, R., Cervantes, J., 2001. A 3-D Lagrangian stochastic model

for the meso-scale atmospheric dispersion applications. Nuclear EngineeringDesign, 208, 15-28.

Riddle, A. Carruthers D., Sharpe, A., McHugh C., Stocker, J., 2004. Compar-ison between FLUENT and ADMS for atmospheric dispersion modeling.Atmospheric Environment, 38, 1029-1038.

Di Sabatino, S., Buccolieri, R., Pulvirenti, B., Britter R., 2007. Simulations ofpollutant dispersion within idealised urban-type geometries with CFD andintegral models, Atmospheric Environment, 41(37), 8316-8329

Scarano, F., 1997. Project Report:Improvements in PIV image processing ap-plication to a backward-facing step. Sint-Genesius-Rode: von Karman In-stitute for Fluid Dynamics.

Spalding, D. B., 1971. Concentration fluctuations in a round turbulent freejet. Journal of Chemical Engineering Science, 26, 95-107.

Snyder, W. H., 1981. Guideline for Fluid Modeling of Atmospheric Diffusion.US Environmental Protection Agency Report EPA-600/8-81-009 pp. 199.

Snyder, W. H., Castro, I.P., 1998. Yaw-response of hot-wire probes at ultra-low

21

Page 22: Stack Gas Dispersion Measurements with Large Scale-PIV

wind speeds. Measurement Science and Technology, 9,1531-1536.Tominaga, Y., Stathopoulos, T., 2007. Turbulent Schmidt numbers for CFD

analysis with various types of flow field. Atmospheric Environment, 41(37),8091-8099.

Tominaga, Y., Mochida, A., Yoshie, R., Kataoka, H., Nozu, T., Yoshikawa,M., Shirasawa., T., 2008. AIJ guidelines for practical applications of CFDto pedestrian wind environment around buildings. Journal of Wind Engi-neering and Industrial Aerodynamics, 96, 1749-1761.

Turner, D. B., 1970. Workbook of atmospheric dispersion estimates. Environ.Prot. Agency, Environ. Health Series Air Pollut., 84.

Wilson, D. J., Lamb, B., 1994. Dispersion of exhaust gases from roof levelstacks and vents on a laboratory building. Atmospheric Environment,28(19), 3099-3111.

van de Hulst, H. C., 1957. Light Scattering by small particles. New York: JohnWiley & Sons.

22