SL-08-005 Final

download SL-08-005 Final

of 12

Transcript of SL-08-005 Final

  • 8/2/2019 SL-08-005 Final

    1/12

    !2008 ASHRAE 45

    ABSTRACT

    A detailed Computational Fluid Dynamics (CFD) study

    of the flow around a Computer Simulated Person (CSP) in anotherwise empty displacement ventilated room is presented.

    Both Reynolds Averaged Navier-Stokes (RANS) and Large

    Eddy Simulation (LES) methods are included in this study.

    The study identifies the requirements of several computational

    aspects that are needed for accurate CFD simulations of the

    personal micro-environment, which include issues related to

    grid size and iterative convergence monitoring, turbulence

    modeling, and radiation modeling. Using the benchmark test

    for evaluating CFD in the indoor environment of Nielsen et al.

    (2003), very good agreement between CFD results and test

    data was obtained using the standard with wall treatment

    and a reduced-order radiation model.

    INTRODUCTION

    For ventilation systems other than an ideal mixing

    system, e.g. displacement and personal ventilation systems,

    the well-mixed assumption can not be applied because spatial

    gradients of the flow, temperature and pollutant fields can be

    large in the vicinity of the Personal Micro Environment

    (PME). Here, the PME is the region around a person that

    affects the air he/she breathes. In a displacement ventilation

    system, human exposure to pollutants is influenced by the

    convective and diffusive transport mechanisms found in the

    thermal plume around the person. The flow in the thermal

    plume is dominated by the buoyancy forces arising from the

    higher temperature of the human body, and this flow field is

    rather complex even in a situation where the person is standing

    in an empty room (Clark and Edholm 1985).

    The PME and the surrounding environment have a sophis-

    ticated relationship that is rarely amenable to simple analytical

    models. Analytical methods are not without merit as they are

    powerful tools for understanding the fundamental features(Awbi 1991) of a real scenario but the interactions of these

    features are not trivial. This concept is elucidated by Melikov

    and Kaczmarcyzk (2007) who state, The measurement at a

    point in a room, without a person present, would not define

    accurately the quality of the air that the person would inhale

    when present at this location. As a result, sophisticated tools

    and methods are needed to extract a deeper understanding of

    the PME.

    Beyond analytical, the two remaining approaches are

    experimental and computational. Experiment has traditionally

    been the most common and reliable method to study the indoor

    environment. The experimental approach could be character-ized by two distinct strategies: field study and detailed

    measurement. Field study involves placing sensors in a rela-

    tively non-intrusive manner within a real environment, statis-

    tically analyze the data obtained and making conclusions

    based on the analysis (Ferro et al. 2003). While this strategy

    provides the most accurate representation of the actual

    scenario, the lack of control of input variables makes quanti-

    tative inference a dubious task. The alternative experimental

    strategy is detailed measurements. Again, sensors are used to

    obtained information about the pertinent quantities, but here

    the experiment is carried out in a more controlled manner with

    higher resolution equipment. For the indoor environment and

    the PME, the types of detailed experimental equipment vary

    widely depending on the desired quantity. For flow velocity,

    common techniques are hot-wire, hot-sphere, particle image

    velocimetry (PIV) and laser-doppler anemometry (LDA), in

    k-"

    Verification and Validation of CFD for thePersonal Micro-Environment

    Chris N. Sideroff, PhD Thong Q. Dang, PhD

    Chris N. Sideroffis a technical sales engineer at Pointwise, Inc., Fort Worth, TX. Thong Q.Dang is a professor of mechanical and aerospace

    engineering, Syracuse University, Syracuse, NY.

    SL-08-005

    2008, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. (www.ashrae.org).

    Published in ASHRAE Transactions Vol. 114, Part 2. For personal use only. Additional reproduction, distribution, or transmission

    in either print or digital form is not permitted without ASHRAEs prior written permission.

  • 8/2/2019 SL-08-005 Final

    2/12

    46 ASHRAE Transactions

    order of increasing accuracy and sophistication. For tempera-

    ture, common techniques are thermal couples, thermistors and

    infra-red imaging. The drawback of such experimental

    systems is that they can be expensive, complex to operate and

    have other limits. Recent examples of detailed experiments

    with thermal manikins in the indoor environment are Bjrn

    and Nielsen (2002) to study the interaction of multiple people,Melikov and Kaczmarcyzk (2007) to examine personal venti-

    lation devices on the PME, Cermak et al. (2006) to identify

    transmission of infectious agents between occupants and Marr

    (2007) to investigate the influence of body motion.

    With the advent of high performance computers at

    commodity prices, computational tools are becoming ever

    more popular. More specifically, Computational Fluid

    Dynamics or CFD has been used with success in a variety of

    areas in the indoor environment community; office buildings

    (Cheong et al. 2003), homes (Huang et al. 2004), hospitals

    (Brohus et al. 2006) and aircraft (Zhang and Chen 2001) are

    examples of a few of these areas. To be able to use CFD as a

    design tool, one first must have confidence that it can predict

    the desired quantities with a certain degree of accuracy. To do

    this, two important questions need to be answered: First, is

    CFD capable of predicting the flow in question and second,

    what is needed to do so? Flows around humans in an indoor

    environment can be particularly difficult to predict because of

    the complex interaction between many different factors. To

    properly answer these questions, standard or canonical bench-

    mark cases are needed so the individual issues may be identi-

    fied along with providing a consistent approach for others

    researchers to use.

    The collaborative efforts of Nielsen et al. (2003) have

    culminated in a benchmark test for evaluating CFD in the

    indoor environment. They have proposed two canonical build-

    ing environment scenarios: mixing and displacement venti-

    lated rooms with a centrally situated manikin. Results and

    discussion of the displacement ventilation case are presented

    in this work. Verification and validation of CFD for the

    personal micro-environment (PME) is necessary for it to

    become a reliable tool. The objective using the benchmark

    case is to identify the key requirements needed to achieve an

    accurate prediction of the detailed flow field around the PME

    of a lifelike Computer Simulated Person (CSP) or manikin.

    Previous studies of Topp (2002) and Topp et al. (2002) have

    examined the differences of simplified and detailed CSP. Theyconcluded that while variations between the two were incon-

    siderable at some distance from the CSP, very near the CSP,

    and more importantly in the personal micro-environment, the

    distinctions were much more apparent. They provided exam-

    ples highlighting how CSP details can affect calculations of

    contaminant transport, heat-transfer coefficients and view

    factors need for radiative heat-transfer. Specifically, Topp

    (2002) provided conclusive evidence that the contaminant

    distribution in the personal micro-environment of the manikin

    is a strong function of geometry detail.

    The CSP geometry used for this study is a digitization of

    a female manikin in the standing position. The data file was

    obtained from Katos research group at the University of

    Tokyo. The surface area of the manikin was approximately

    1.48 m2, a value within the known measured range for females.

    Diminutive body details such as ears, fingers and toes were

    removed to disburden grid generation while maintaining asufficient level of surface description. Furthermore, aspects

    such as hair and clothes were also not included to alleviate

    further uncertainties (e.g., heat transfer through clothing).

    Therefore the manikin represents a female of average body

    surface area situated in the standing posture.

    The original summary of the scenario setup, boundary

    conditions and suggested reporting method of the two bench-

    mark cases presented in this paper are outlined in Nielsen et al.

    (2003). CFD results were extracted and compared at the same

    locations where experimental data was measured. The exper-

    imental data, in addition to further information, for each case

    is openly available at http://www.cfd-benchmarks.com.

    COMPUTATIONAL MODEL

    All CFD simulations were performed using the commer-

    cial CFD software FLUENT (version 6.3), where the incom-

    pressible Navier-Stokes equations (Equation 1) are solved

    with the SIMPLEC pressure-velocity coupling method along

    with the energy equation (Equation 2).

    (1)

    (2)

    The overbar represents averaged quantities whereas

    primed variables are fluctuating quantities. The second last

    term in Equation 1 is the Boussinesq approximation for ther-

    mal buoyancy effects and the last terms in Equations 1 and 2

    are the Reynolds stress tensor and heat-flux vector, respec-

    tively. To close the problem the Reynolds stress tensor and

    heat-flux vector are related to the mean velocities and temper-

    atures through the gradient-diffusion hypothesis

    (3)

    (4)

    where is the mean strain-rate tensor and

    is the turbulence kinetic energy. The turbulent viscosity, , is

    determined from characteristic velocity and length or times

    scales.

    (5)

    D v

    Dt--------

    1

    #o----- P$ $+ %$ v& ' g 1 () T& 'k $ v 'v '& '**=

    DT

    Dt-------- $ +$ T& ' $ T'v'& '**=

    v 'v '& ' %TS2

    3---kI+=

    T'v'& ' +T$T=

    S1

    2--- $ v $ v

    T

    +, -. /= k

    %T

    %T C0VL ; %T C0V2T==

    2008, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. (www.ashrae.org).

    Published in ASHRAE Transactions Vol. 114, Part 2. For personal use only. Additional reproduction, distribution, or transmission

    in either print or digital form is not permitted without ASHRAEs prior written permission.

  • 8/2/2019 SL-08-005 Final

    3/12

    ASHRAE Transactions 47

    The turbulent thermal diffusivity, and

    . The specification of the characteristic scales are

    computed by the various turbulence models.

    The numerical convective scheme used for all transport

    equations was a 2nd order accurate upwind scheme, while

    diffusion terms are discretized using the 2nd order accurate

    central difference scheme. Pressure interpolation wasachieved with a 2nd order accurate scheme similar to the 2nd

    order upwind convective scheme. Computational grids were

    generated using the commercial grid-generation software

    Gridgen. A Beowulf cluster with 64 1.6 GHz processors (x86-

    64 architecture) and 64 Gbytes of total system memory made

    this type of analysis possible. Thirty-two processors were

    typically used, which took anywhere from 48 to 72 hours

    depending on the grid resolution and turbulence models, the

    latter include the model with enhanced wall treatment and

    the model for the Reynolds-Averaged Navier Stokes

    (RANS) equations, and the Large Eddy Simulation (LES)

    method.

    In displacement ventilation systems, clean and cool air is

    introduced into the lower portion of the room at a low velocity

    so not to cause draught discomfort. The cool supply air then

    spreads throughout the occupied zone, i.e. the zone occupied

    by people. Heat sources present in the room cause the air to

    rise due to thermal buoyancy. As the warming air rises,

    contaminants picked up along the way are carried above the

    occupied zone. The warm, dirty air is then exhausted in the

    upper portion of the room. Figure 1 illustrates the benchmark

    displacement ventilation setup used (Nielson et al. 2003). The

    room has dimensions of 2.5 m wide, 3.5 m deep and 3.0 m high

    (width, depth and height correspond to the x,y andz coordi-

    nates respectively). Air is supplied through a 0.4 m wide by 0.2m high rectangular hole located on the floor and is discharged

    through a hole of equivalent shape and size at the ceiling on the

    opposite wall. Measured from the top of the head, the CSP is

    located 1.75 m downstream of the inlet wall, centered in thex-

    direction and 1.75 m from the floor. Experimental data

    included PIV data measured in three windows. These three

    windows are the small gray rectangles indicated in Fig. 1.

    Within each window, velocity components in the plane of

    measurement were provided along a 0.2 m long line. Theselines are labeled L1, above the head, L2, projecting from

    the center of the mouth and L3, projecting from the center of

    the torso. All three horizontal locations are in thex mid-plane.

    These test data are used for CFD validation.

    In the CFD calculations, the computational domain

    consists of the entire room shown in Figure 1. Several compu-

    tational grids of different mesh size and topology are gener-

    ated in this work (to be discussed in details later). The

    following boundary conditions were used for all simulations.

    The velocity and temperature at the inlet were 0.2 m/s and

    22C respectively. A turbulence intensity of 30% and length

    scale of 0.1 m were experimental values suggested by Nielsen

    et al (2003). The only variable specified at the outlet was static

    pressure and was set to zero. The no-slip condition along with

    zero heat-flux (insulated) was applied to all the room walls.

    Along with the no-slip condition on the remaining bound-

    ary, the manikin surface, a boundary condition for the energy

    equation is required that approximates a human body. The

    human body is a dynamic thermoregulatory system (Fiala et

    al. 1999; Tanabe et al. 2002). It has the ability to adjust heat

    output in response to global and local changes in skin temper-

    ature. The objective of this work was not to predict the thermal

    response of a human but to evaluate the ability of CFD to

    predict the flow in the personal micro-environment. As a

    result, a thermoregulation model was not used to control the

    heat output of the body. Rather, a constant (spatially and

    temporally) heat-flux was chosen to mimic the heat exchange

    of the body with its surroundings. The net heat loss by a human

    is not only affected by their local environment but also can

    vary widely depending on gender, age and physiological

    makeup. The manikin used in this work represents a young

    adult female with a BMI in the normal range1. A characteristic

    heat loss for a human of this makeup is around 76 W. Using

    this value along with the surface of area of the manikin (1.48

    m2), a heat-flux of 51.2 W/m2 is obtained. This value is the

    total heat-flux per unit area, where the fraction due to convec-

    tion and radiation are determined during the calculation. It iswidely accepted that roughly half of human heat loss is due to

    convective means while the remainder is due to radiation. An

    often used approach is to ignore radiation to avoid the

    perceived difficulties associated with radiation calculations.

    For cases in which only convection was modeled, half of the

    aforementioned heat-flux per unit area value was assumed

    (25.6 W/m2).

    +T %T PrT1=PrT 0.85=

    k-"%2-f

    1. BMI, body-mass index, is a measure of body fat based on heightand weight.Figure 1 Displacement ventilation room configuration.

    2008, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. (www.ashrae.org).

    Published in ASHRAE Transactions Vol. 114, Part 2. For personal use only. Additional reproduction, distribution, or transmission

    in either print or digital form is not permitted without ASHRAEs prior written permission.

  • 8/2/2019 SL-08-005 Final

    4/12

    48 ASHRAE Transactions

    VERIFICATION

    Following the definitions of Roache (1997), verification

    can be thought of assolving the equations rightwhereas vali-

    dation would be solving the right equations. Verification

    involves quantifying the error induced by solving the chosen

    equations using discrete approximations. The total error

    includes those due to coding, discretization, grid convergenceand iterative convergence.

    Because a commercial CFD code was used and access to

    the source code is not possible, code verification could not be

    performed. While, the author recognizes this as a potential

    source of error, due to the large FLUENT user base, it is

    assumed that any unreported coding errors negligibly affect

    the solution. A list of known resolved and unresolved errors in

    FLUENT are available on their supported user website. Solu-

    tions to the equations were computed using a well-docu-

    mented second-order discretization scheme but no verification

    of the observed order of accuracy was performed. The author

    recognizes this as a potential source of error but believes it tobe reasonable assumption again due to large FLUENT user

    base that accept the implementation of this discretization

    scheme to be second-order or close to second-order.

    The verification of FLUENT performed in this work

    involves grid and iterative convergence studies. Grid conver-

    gence will be demonstrated by comparing relevant quantities

    on the sequence of grids at strategically chosen locations. A

    more rigorous approach to determine grid convergence is the

    Richardson Extrapolation (RE) method (Richardson and

    Gaunt 1927). Two requirements of RE are constant grid refine-

    ment ratio and values at the same location. While the later is

    possible through interpolation, the former is not for the typesof grids used in this work. Furthermore, the reliability of these

    methods on complex, unstructured grids remains unclear. As

    such error estimation using the RE method was not used. Iter-

    ative convergence for all solutions was performed and an

    example highlighting the strategies of monitoring conver-

    gence will be discussed.

    Grid Convergence Study

    Due to the unknown behavior of the global and local flow

    features in both cases, determining the appropriate grid strat-

    egy was non-trivial. Traditionally the number of grid cells has

    not exceeded several hundred thousands, due in large extent tolack of computational resources (Posner et al. 2003; Xing et al.

    2001; Murakami 2004; Zhu et al. 2005; Srebric et al. 2007). In

    the present study, grids consisting of up to seven million cells

    were investigated. However, even with a seemingly indispens-

    able amount of processing power, care should be exercised

    when determining the global resolution (total cells), local

    resolution (clustering around the CSP) and topology (cell

    type) for the grids in these scenarios.

    The approach to creating each grid began with meshing

    the surface of the manikin. A consequence of the complex

    manikin surface was a non-uniform distribution of triangles.

    This was done to ensure all the details of the geometry were

    represented. Next, the walls, inlets and outlets were meshed

    uniformly. Finally, the volume grid was created where a

    growth rate factor (or the rate at which the cell size increases)

    of 1.2 or less was used to gradually increase the cell size away

    from the manikin. Some grids included an additional stepwhere several layers of prismatic cells on the manikin were

    created to sufficiently resolve the boundary layer (called

    boundary layer grid here). In all cases, the equi-volume skew-

    ness values of the interior cells (or the quality of the cells) were

    monitored to ensure that they fall within the guidelines of the

    solver FLUENT.

    To establish grid independent solutions for the displace-

    ment ventilation case four grids were used. A summary of the

    four grids is provided in Table 1. The four grids consisted of

    two with strictly tetrahedral cells and two with the boundary

    layer cells added near the manikin. The important parameters

    that differentiate these grids include the number of triangles

    used to define the manikin geometry, the average value on

    the manikin surfaces2, and the type of grid (with or without

    boundary layer prismatic cells). Recall that the growth factor

    of the cell volume was kept under 1.2 and grid quality as

    measured by the equi-volume skewness was within the recom-

    mended value of the CFD solver. Illustrations of each grid type

    can be found in Sideroff (2007).

    The solutions from each grid used for the grid dependency

    study were obtained using the standard turbulence model

    with enhanced wall treatment, and only the convective portion

    of heat-transfer. Since convection alone was included, only

    half the heat-flux value was applied to the manikin surface.

    The three measurement stations (head, face, and torso)

    detailed in Figure 1 are used to compare solutions from the

    four grids. Besides being located where the data were

    measured, these stations are meaningful because they pene-

    trate the thermal plume at three distinctly different locations.

    Comparisons of the vertical velocity were made at the

    three stations illustrated in Figure 1 for the four grids, and the

    largest differences between these solutions occurred at the

    torso (station L3). These results are illustrated in Figure 2,

    which show that Grid C and Grid D appear to exhibit grid inde-

    pendence while Grid A and Grid B do not. In particular, the

    steep gradient of the near-wall velocity is difficult to capture

    when a strictly tetrahedral grid is used, unless of courseextremely small tetrahedrons are used. Employing strictly

    tetrahedral to resolve the boundary layer would result in grids

    well in excess of ten million cells.

    2. Where is the distance from wall adjacent cell-

    center to the wall, the friction velocity and is the

    kinematic viscosity.

    y+

    y+ypu

    2

    v----------- : yp=

    u23wall

    #-----------= %

    k-"

    2008, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. (www.ashrae.org).

    Published in ASHRAE Transactions Vol. 114, Part 2. For personal use only. Additional reproduction, distribution, or transmission

    in either print or digital form is not permitted without ASHRAEs prior written permission.

  • 8/2/2019 SL-08-005 Final

    5/12

    ASHRAE Transactions 49

    Iterative Convergence Study

    When steady-state RANS simulations are used, conver-

    gence is typically done by monitoring the average residualsduring the iterative process of the transport equations (mass,

    momentum energy, and turbulence model equations). Resid-

    uals were monitored and typically dropped four orders of

    magnitude for the continuity equation (two to three orders

    lower for other equations). The number of iterations required

    to reach convergence varied but was at least several thousands.

    For the test case studied here, it was noted that average resid-

    uals alone were not good indicators of convergence. Averag-

    ing of the residuals can hide regions where the local residual

    may be many orders of magnitude larger. It was found that

    along with residuals, velocity magnitude should be monitored

    at strategically chosen points to give a more accurate indica-

    tion that the solution had indeed converged. For the presentproblem, two points were chosen: one in the thermal plume

    (referred to as point A), and another away from and behind the

    CSP to monitor the room airflow (referred to as point B).

    During the iteration process, two distinct flow structures

    developed: the thermal plume, which was expected, and

    another not so obvious structure was a rather complex re-

    circulating flow around the room and behind the CSP gener-

    ated by the inlet vent. The momentum of the inlet vent created

    a low-speed jet that progressed along the floor, and instead of

    rising upon hitting the warm feet of the CSP, it proceeded

    along to the back wall. In fact, the flow circulated several times

    behind the CSP, three or more times in some instances, before

    being entrained in the thermal plume. The development of

    these two flow structures was not independent of each other

    thus making judgment of convergence non-trivial.

    As mentioned, FLUENT uses averaged residuals to moni-

    tor convergence of the transport equations. In the present prob-

    lem, the continuity residual typically dropped about four

    orders of magnitude and leveled out after about 10,000 itera-

    tions, and in this case, falsely indicating convergence. It is

    noted that, since the velocity magnitude is highest in the ther-

    mal plume, the average residual is basically a convergence

    indicator of the flow structure in the thermal plume. When the

    velocity magnitudes at Points A and B were monitored during

    the iteration process, it was observed that convergence was not

    yet reached after 10,000 iterations. Figure 3 shows the velocity

    in the plume (Point A) declines monotonically, stops changing

    at about iteration 9,000, and remained constant until about

    iteration 10,000 where it gradually increases back to about

    0.035 m/s from a value of 0.01 m/s. At iteration 14,000, the

    velocity magnitude began to slowly decline again and eventu-

    ally leveling out to a value of 0.02 m/s at roughly iteration

    27,000. Along the floor (Point B), the velocity magnitude

    made a couple of large oscillations during the first 7,000 iter-

    ations and then began to oscillate at a much higher frequency

    but at a constant mean value. Eventually the oscillations died

    away and the velocity became steady near iteration 27,000.

    During the first several thousands iterations (

  • 8/2/2019 SL-08-005 Final

    6/12

    50 ASHRAE Transactions

    be very large if the sources of contaminants are located behind

    the CSP in the vicinity of the re-circulating flows.

    VALIDATION

    It is validation when one wants to determine how well the

    computational model can represent reality. Through verifica-

    tion, one gains confidence that the CFD code can produce

    solutions to the discrete equations of known error but provides

    no guarantee these equations sufficiently captures the physics

    of the problem. To validate CFD codes, experimental data

    must be used because experiments themselves do not selec-

    tively exclude the relevant physics. When validating a compu-

    tational model, the user must now account for not only the

    error associated with the computational model but also theerror associated with the experiment. Therefore error due to

    measurement uncertainty and error to facility biasing should

    be also accounted for.

    Katos research group at the University of Tokyo has

    obtained experimental data of the displacement ventilation

    case studied here. The setup of Katos experiment is similar to

    the computational model shown in Figure 1 where the manikin

    is suspended slightly above the floor. Two velocities compo-

    nents were measured with a particle image velocimetry (PIV)

    system near the manikin and values for each were provided

    along the lines indicated in Figure 1. Velocity magnitude,

    turbulent intensity and temperature were measured at several

    locations away from the manikin including the inlet. Unfortu-nately no estimates of uncertainty or error were provided.

    Turbulence Modeling

    Due to our inability to compute the entire range of flow

    scales for all but the most trivial problems, modeling of turbu-

    lence has become a research science of its own. The most ubiq-

    uitous approach to modeling turbulence is the RANS method

    while the more robust but computationally demanding

    approach is the LES method. Two RANS models as well as an

    LES model were investigated in this work.

    Eddy-viscosity RANS model are the most popular class

    of turbulence model utilized for indoor environment simula-

    tions. They are attractive because they offer the best balance

    between accuracy, complexity and computational cost. The

    most popular of these, the standard (Jones and Launder

    1972), was used in this work. The transport equations for the

    turbulence kinetic energy, , and turbulence dissipation rate,

    , respectively are as follows:

    (6)

    (7)

    where is the turbulent viscosity, is

    thermal buoyancy induced production and is the time

    scale. Complete details of these equations, including the

    model constants, can be found in many other books on turbu-lence (Pope 2000; Tennekes and Lumley 1972).

    To extend the applicability of the standard model,

    FLUENT has made available an enhanced wall treatment

    option. The enhanced wall function is a near-wall modeling

    approach that enables the standard equations to be inte-

    grated all the way to the wall. Essentially, this provides the

    standard with the ability to resolve the boundary layer

    through the buffer layer into the viscosity affected sub-layer

    without the need for an explicit wall function. To allow this, a

    two-layer model is used to specify and in the viscous sub-

    layer. Providing the standard with the ability to resolve the

    boundary layer imposes requirements on , i.e. less than

    one. Complete details of the enhanced wall treatment are

    found in the FLUENT manual.

    A state-of-the-art RANS model, dubbed the , with

    some noteworthy improvements over the standard , was

    also used in this work. Regardless of the ability to full resolve

    the boundary layer, it is well known that the standard

    model over-predicts the production of turbulent kinetic energy

    near solid walls. The performance of the definition of the

    eddy viscosity degrades rapidly as solid walls are approached

    primarily due to incorrect scaling of turbulent kinetic energy.

    The model of Durbin (1991) changes the definition of the

    velocity scale to primarily because scales as

    as it approaches solid walls whereas scales as (Durbin1991); thus providing stronger damping as the wall is

    approached. Furthermore, an elliptic relaxation function, , is

    included to account for the non-local damping effects of solid

    boundaries. Through this approach the elliptic relaxation

    function can account for anisotropy present near solid bound-

    aries. There have been various modifications to the model

    and the versions of the and equations incorporated in

    FLUENT are identical to Model 3 of Sveningsson (2003).

    Along with the Equations 6 and 7, the remaining equations that

    constitute this version of the are as follows:

    Figure 3 Convergence history of velocity at Points A and B.

    k-"

    k

    "

    Dk

    Dt------- $ %

    %T4k-----+, -

    . / $k, -. / 2%TS

    2 " Pb+ + +*=

    D"Dt------- $ %

    %T4"-----+, -

    . / $", -. / C1"

    T-------- 2%TS

    2 C3" Pb+& ' C2""T---+*=

    %T C0k

    "--= Pb g+T(

    5T5z------=

    T k"--=

    k-"

    k-"

    k-"

    " %tk-"

    y+ y+

    %2-fk-"

    k-"

    k-"

    %2-fv'2& '1 21

    2v' y4

    k y2

    %2-fv2

    %2-f

    2008, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. (www.ashrae.org).

    Published in ASHRAE Transactions Vol. 114, Part 2. For personal use only. Additional reproduction, distribution, or transmission

    in either print or digital form is not permitted without ASHRAEs prior written permission.

  • 8/2/2019 SL-08-005 Final

    7/12

    ASHRAE Transactions 51

    (8)

    (9)

    where is the turbulent viscosity. The time scale,T, defined to allow adjustment near walls and model constants

    can be found in Sveningsson (2003).

    While the standard employs ad-hoc wall treatments,

    the model instead handles the near-wall behavior through

    a more physical approach: the and equations. Without the

    need for wall functions, the turbulence model requires

    that the boundary layer be fully resolved and should be less

    than one.

    Eddy-viscosity RANS models, such as the standard

    and , are known to be deficient for low Reynolds

    numbers flow. The majority of the assumptions on which

    they are based can be traced to the fully turbulent approxi-

    mation. Furthermore, turbulence production from thermal

    gradients is a particularly challenging issue for typical RANS

    models. While there has been considerable effort to extend

    the capabilities of traditional RANS models (Murakami et al.

    1996; Kenjere et al. 2002, 2005), these new models tend to

    be overly complex requiring additional equations, more coef-

    ficients and typically introduce numerical instability. Despite

    their achievements, these models undoubtedly fall well short

    of being reliable for a wide range of indoor environment

    flows. With this in mind, it seems reasonable to consider

    models beyond the RANS approach. LES is a more robust

    method of approximating complex turbulent flows than

    RANS methods. In contrast to RANS modeling, Large EddySimulation (LES) computes the large-scale motions of the

    flow directly. The small-scale, dissipative motions of turbu-

    lence tend to more amenable to modeling because of their

    more uniform character, whereas the large-scale motions

    contain the majority of the energy and anisotropy. As a result,

    LES is expected to be more accurate, particularly in complex

    flows where the assumptions inherent to RANS models

    rarely exist. The drawback is that LES simulations are always

    three dimensional, unsteady one. Provided the boundary

    layer is resolved and y+ is less than one, a linear stress-strain

    relationship is assumed for LES. If however the first cell does

    not lie well within the viscous sub-layer, it is assumed the

    first cell lies within the logarithmic layer and the traditionallaw-of-the-wall is used.

    LES typically does not suffer the drawbacks associated

    with RANS models provided the energy-containing length

    scales are sufficiently resolved. The question is how to ensure

    this happens without a priori knowledge of the flow. Because

    LES computes the large scale motions and models the small

    scale ones, knowledge or even approximation of the small

    scales would be beneficial. From known values of the length

    and time scales, estimates of the smallest scales, the Kolmog-

    orov scales (Tennekes and Lumley 1972), can be made. The

    Kolmogorov length ( ) and time ( ) scales are be expressed

    in terms of the kinematic viscosity, %.

    (10)

    In determining and , only an estimate for e remains.

    Following Tennekes and Lumley (1972), an inviscid estimatefor can made through the relationship . The fluc-

    tuating velocity, , is determined by the relation .

    Here, would be the peak plume velocity, 0.3 m/s. The turbu-

    lent intensity, , and turbulent length scale m

    of the plume were obtained from Marr and Glauser (2006) and

    Marr (2007). Using these values, m2/s3. From this,

    the estimated Kolmogorov scales of the thermal plume are

    mm and .

    Now that approximations of the smallest scales are avail-

    able, determination of the appropriate grid spacing and time

    step for LES can be made. If a DNS of this flow was

    performed, a grid spacing and time step would be chosen to

    match the Kolmogorov values. However, since LES models

    the small scales a larger grid spacing and time step can be used.

    Meyers et al. (2003) recommends the LES sub-grid scale

    cutoff, i.e., grid spacing, to be around 1020 Kolmogorov

    scales. This would suggest a grid spacing of about 6 mm would

    be sufficient for LES. The time step would be determined in a

    similar manner but numerical concerns must also be consid-

    ered. The recommendation of Meyers et al. (2003) applied to

    the time step would yield 0.25 s but the flow time for a 6 mm

    cell with a velocity of 0.3 m/s would be 0.02 san order of

    magnitude smaller than . It is unlikely that the maximum

    velocity would occur in the smallest cell, since the smallest

    cells are located adjacent that walls. The appropriate time stepsize would lie somewhere between 0.02 and 0.25.

    The dynamic Smagorinsky sub-grid scale model of

    Germano et al. (1991) was used in this work. The sub-grid

    scale viscosity and length scale are defined as

    (11)

    (12)

    where the over-tilde denotes spatial average (as opposed to

    temporal averaging for RANS models) and the filter width,

    . Complete details along with the model constants

    can be found in Germano et al. (1991).As previously mentioned, results from the turbulence

    models were compared with the experimental data at the loca-

    tions indicated in Figure 1. Upon examining the vertical veloc-

    ity profiles in Figure 4 above the head, significant differences

    are revealed between not only the turbulence models but also the

    experimental values. The general profile shapes appear reason-

    able, but only a small the portion of the profile near the head

    (

  • 8/2/2019 SL-08-005 Final

    8/12

    52 ASHRAE Transactions

    by 0.1 m/s. It is evident from Figure 4 that LES provides

    marginal improvement. Near the head, the LES profile is consis-

    tent with the slope of the data. Beyond the near-wall region, theLES profile compares better to the test data than either RANS

    models, but the improvement is marginal. The LES results over-

    predict the magnitude by as much as 0.05 m/s.

    At the face shown in Figure 5, the RANS results yielded

    considerably different trends. Overall, the model under-

    predicts the thickness of the thermal plume, while the

    model over-predicts it. Near the wall (

  • 8/2/2019 SL-08-005 Final

    9/12

    ASHRAE Transactions 53

    side walls and floor without radiation are several degrees lower

    than with radiation, while the converse is true on the ceiling.

    Indubitably, the incorporation of radiation has consequences

    on not only on the individual contribution of heat transfer

    modes but also on the surface temperatures.

    To evaluate the effect of radiation more definitively,

    results were compared to the test data at the three locationsindicated in Figure 1. Upon inspection of the results of Figure

    9 above the head, it is immediately obvious the influence radi-

    ation has on the thermal plume. The over-prediction of the

    magnitude seen in Figure 4 without radiation appears to be

    remedied by the inclusion of radiation. The LES profile is in

    excellent agreement with the data, and the standard and

    yield more than satisfactory results.

    The results at the face with radiation included shown in

    Figure 10. While not in agreement to the degree seen above the

    head, these results still provide improvement over predictions

    without radiation included. Here, it could be argued the LES

    provides no considerable benefit over the standard or

    . All numerical profiles show roughly the same peak value

    albeit higher than the data0.125 m/s for the , 0.11 m/s

    for LES and standard versus 0.08 m/s for the data.

    However, beyond about 0.03 m the numerical profiles follow

    the data remarkably close.

    The final location at the torso indicates similar trends as

    seen at the face. As shown in Figure 11, the peak value is over-

    predicted by both models but beyond there they follow the data

    reasonably well. The standard does not predict the change

    in slope seen in the data or the and LES but overall yields

    acceptable results.

    Reduced-Order Model for Radiation

    The calculations with radiation provided evidence that

    radiation modeling cannot be ignored, but the inclusion of

    radiation modeling did not come without a cost. For surface-

    to-surface radiation calculations, the view factors are needed.

    Although purely a pre-processing step, computing the view

    factors can be computationally expensive (Sideroff 2007). It

    would be beneficial if the effects of radiation could be

    included without actually performing the radiation modeling.

    If the surface temperatures were somehow known a priori, the

    effects of radiation can be included without the difficulties of

    actually computing radiative energy transfer. Unfortunately,

    determining a detailed spatial (or temporal) description of the

    surface temperature from experiment is exceedingly difficult

    or simply not possible, nor it is the case in practice.

    It was postulated that perhaps a reduced-order or low-

    order description of the surface temperature would be

    adequate. And if a low-order description of the surfacetemperature were used, how would this affect the flow in the

    PME? To answer this question, an examination of the impact

    of a reduced surface temperature description was carried out.

    The simplest way is to reduce the surface temperature of each

    surface to an average value.

    Figure 6 Vertical velocity at torso (station L3).

    Figure 8 Surface temperatures with radiation; LES model

    on grid D.

    k-"%2-f

    k-"%2-f

    %2-fk-"

    k-"%2-f

    Figure 7 Surface temperatures without radiation; LES

    model on grid D.

    2008, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. (www.ashrae.org).

    Published in ASHRAE Transactions Vol. 114, Part 2. For personal use only. Additional reproduction, distribution, or transmission

    in either print or digital form is not permitted without ASHRAEs prior written permission.

  • 8/2/2019 SL-08-005 Final

    10/12

    54 ASHRAE Transactions

    Using the solution from the standard on grid D, an

    area-weighted average of the temperature distribution was

    computed for each surface. The resulting average surface

    temperatures and those subsequently used for boundary

    conditions are tabulated in Table 2.

    A calculation using the standard on grid D was

    performed using the average surface temperatures from

    Table 2 and compared against the results from the standard

    with the unaltered surface temperatures from radiationmodeling, along with the test data. Comparing the solutions at

    the three stations where test data are available, it was found

    that excellent agreement were obtained at station L1 (above

    the head) and station L2 (at face). At station L3 or the torso

    shown in Figure 12, more noticeable differences are observed.

    The peak value of the profile with the reduced-order model is

    marginally lower than with the full radiation modeling

    included. However, the shapes of the profiles are very similar.

    CONCLUSIONS

    A detailed verification and validation study using a

    commercial CFD code (FLUENT) of the flow around a

    Computer Simulated Person in a displacement ventilation

    room was carried out. Following the guidelines of the bench-

    mark displacement ventilation case of Nielsen et al. (2003)

    several recommendations concerning the verification and vali-

    Figure 9 Vertical velocity above head with radiation

    (station L1).

    Figure 11 Vertical velocity at torso with radiation

    (station L3).

    Table 2. Average Surface Temperatures

    Surface Average Temperature (K)

    manikin 303.9

    floor 297.9

    ceiling 298.5

    side walls 298.3

    front wall 298.3rear wall 298.4

    k-"

    k-"

    k-"

    Figure 10 Vertical velocity at face with radiation

    (station L2).

    Figure 12 Vertical velocity at torso comparing radiation

    modeling and averaged surface temperatures

    (station L3).

    2008, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. (www.ashrae.org).

    Published in ASHRAE Transactions Vol. 114, Part 2. For personal use only. Additional reproduction, distribution, or transmission

    in either print or digital form is not permitted without ASHRAEs prior written permission.

  • 8/2/2019 SL-08-005 Final

    11/12

    ASHRAE Transactions 55

    dation of CFD for the PME were elucidated by examining grid

    dependency, iterative convergence, turbulence modeling and

    radiation modeling.

    Four grids of increasing resolution and varying topologies

    were constructed to demonstrate grid independence. Due to

    the complex geometry of the CSP, the grids used were of the

    unstructured type. It was found that achieving grid indepen-

    dent solutions while maintaining an acceptable cell countusing a strictly tetrahedral topology was exceedingly difficult.

    The two finest grids included a technique where several layers

    of thin triangular prism cells were created by extruding the

    surface triangles away from the CSP. These layers of prismatic

    cells allowed full resolution of the boundary layer without

    significant increase in the number of cells. To achieve grid

    independent solutions, approximately 100,000 surface trian-

    gles are needed on the CSP, at least several layers of prismatic

    cells around the CSP to achieve values less than one and a

    maximum growth rate of 1.2 for the remaining tetrahedral

    cells around the CSP.

    To achieve convergence for steady-state (RANS) calcu-

    lations, quantities other than average residual needed to bemonitored. Because two distinct flow structures developed,

    averaged residual was not a sufficient convergence parameter.

    Velocities were monitored at strategically chosen pointsone

    in the plume and the other away from the CSP in the recircu-

    lating room flowduring the iteration process. While the

    velocity in the thermal plume appeared to reach steady-state at

    approximately the same number of iterations as the residualsthis was not so. The velocity of the recirculating flow near the

    floor continued to oscillate for several thousand iterations

    beyond the apparent steady state of the plume. Nearly

    30,000 iterations were needed to achieve convergence. It was

    noted that if the velocity in the PME of CSP were of interest,

    the error incurred by incomplete convergence may be incon-sequential. However, if, for example, contaminant transport

    from a source away from the CSP in the recirculating flow

    were of interest, iterative convergence error may be important.

    Due to the perceived effort required to include it, radia-

    tion modeling is commonly avoided by assuming half the

    CSP heat loss for convection only. This approach requires

    using a heat-flux (first derivative) boundary condition.

    Through careful examination of the coupling of surface-to-

    surface radiation equations to the flow equations and results

    that include radiation modeling it was found that neglecting

    radiation modeling when heat-flux boundary conditions are

    used is erroneous. While the actual radiative to convective

    heat transfer ration was closer to 60/40 (not 50/50), it wasconcluded that the significant change in wall temperatures

    due to radiation caused the differences and not the lower

    convective component.

    If somehow the actual surface temperatures were known

    a priori, then the effects of radiation could be included without

    actually including a radiation model. The actual surfacetemperatures would need to be obtained from experiment

    because reality does not disclude radiation. However, obtain-

    ing a detailed description of the surface temperatures experi-

    mentally would be exceptionally difficult or simply not

    possible. By applying spatially averaged values of surface

    temperatureobtained from the calculations with radiation

    includedit was found that velocities in the PME were rela-

    tively insensitive to this approach. This suggests that measur-ing the surface temperature at few points on all the

    participating walls and using those as temperature boundary

    conditions for CFD calculations would yield satisfactory solu-

    tions.

    Turbulence models tend to be the most unreliable aspect

    of CFD, particularly for low Reynolds, thermally buoyant

    flows encountered in the PME. Results from turbulence

    models were validated with high-resolution PIV data suppliedby Katos research group at the University of Tokyo. The stan-

    dard with an enhanced wall-treatment, the model of

    Durbin (1999) and dynamic Smargorinsky LES turbulence

    models were utilized. Without radiation (or the effects)

    included none of the three turbulence models provided satis-

    factory results. Only when radiation (or the effects) were

    included did any turbulence yield reasonable comparison to

    the data. Regardless of the innovative improvements of the

    , it did not yield improvement over the standard . LES

    does not suffer from the drawbacks of RANS models butdespite its wider applicability did not offer any benefit over

    the standard . When an enhanced wall treatment was usedwith a grid that resolves the boundary layer, the standard

    model was found to provide solutions as accurate as the

    or LES.

    ACKNOWLEDGMENTS

    This work was supported by the US Environmental

    Protection Agency (EPA) through the STAR Center for Envi-

    ronmental Quality Systems (www.eqstar.org) and the Center

    of Excellence in Environmental Systems (www.coees.org) atSyracuse University.

    REFERENCES

    Awbi, H.B. 1991. Ventilation of Buildings. E & FN Spon

    Publishing.

    Bjrn, E. and Nielsen, P.V. 2002. Dispersal of exhaled air

    and personal exposure in displacement ventilated rooms.

    Indoor Air 12(2): 147-164.

    Cermak, R. Melikov, A. Forejt, L. and Kovar, 0. 2006. Per-

    formance of personalized ventilation in conjunction

    with mixing and displacement ventilation.International

    Journal of HVAC Research. 12: 295-311.

    Cheong, K.W.D., Djunaedy, E., Pho, T.K., Tham, K.W.,

    Sekhar, S.C., Wong, N.H. and Ullah, M.B. 2003. Mea-

    surements and computations of contaminants distribu-

    tion in an office environment. Building and

    Environment. 38: 135-145.

    Clark, R.P. and Edholm, O.G. 1985. Man and His Thermal

    EnvironmentLondon, UK: Edward Arnold.

    Durbin, P.A., 1991. Near-wall turbulence closure modeling

    without damping functions. Theoretical and Computa-

    tional Fluid Dynamics 3: 1-13.

    Fenske, J.D., Paulson, S.E. 1999. Human breath emissions of

    VOCs.Journal of Air and Waster Management49: 594-

    598.

    y+

    k-" %2-f

    %2-f k-"

    k-"k-"

    %2-f

    2008, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. (www.ashrae.org).

    Published in ASHRAE Transactions Vol. 114, Part 2. For personal use only. Additional reproduction, distribution, or transmission

    in either print or digital form is not permitted without ASHRAEs prior written permission.

  • 8/2/2019 SL-08-005 Final

    12/12

    56 ASHRAE Transactions

    Ferro, A., Kopperud, R.J., and Hildeman, L. 2004. Source

    strengths for indoor human activities that resuspend par-

    ticulate matter. Environmental Science & Technology38(6): 1759-1764.

    Fiala, D., Lomas, K.J. and Stohrer, M. 1999. A computer

    model of human thermoregulation for a wide range of

    environmental conditions: the passive system. Journal

    of Applied Physiology 87(5): 1957-1972.

    Germano, M., Piomelli, U., Moin, P. and Cabot, W.H. 1991.

    A dynamic sub-grid scale eddy viscosity model.Physics

    of Fluids A 3: 1760-1765.

    Huang, J.M., Chen, Q., Ribot, B., and Rivoalen, H. 2004.

    Modeling contaminant exposure in a single-family

    house.Indoor and Built Environment13: 5-19.

    Jones, W.P. and Launder, B.E. 1972. The prediction of lami-

    narization with a two-equation model of turbulence.

    International Journal of Heat and Mass Transfer 15:

    301-314.

    Kenjere, S., Hanjalic, K. and Gunjaro, S.B. 2002. A T-

    RANS/VLES approach to indoor climate simulations.

    Proceedings of ASME 2002 Fluids Engineering Divi-sion Summer Meeting, Montreal, CA.

    Kenjere, S., Gunjaro, S.B. and Hanjalic, K. 2005. Contribu-

    tion to elliptic relaxation modelling of turbulent natural

    and mixed convection. International Journal of Heat

    and Fluid Low 26:. 569-586.

    Marr, D.R. and Glauser, M.N. 2006. Length scale propaga-

    tion along a joint inlet and thermal buoyancy driven

    Flow.AMSE International MeetingPaper 98532.

    Marr, D.R. 2007. Length scale propagation along a joint inlet

    and thermal buoyancy driven flow. PhD Thesis, Syra-

    cuse University, Dept. of Mechanical and Aerospace

    Engineering.

    Melikov, A. and Kaczmarcyzk, J. 2007. Measurement andprediction of indoor air quality using a breathing ther-

    mal manikin.Indoor Air 17: 50-59.

    Meyers, J. Geurts, B.J. and Baelmans, M. 2003. Databaseanalysis of errors in large-eddy simulation. Physics of

    Fluids 15(9): 2740-2755.

    Modest, M.F. 2003. Radiative Heat Transfer. San Diego,

    CA, USA: Academic Press.

    Murakami, S., Kato, S., Chikamoto, T., Laurence, D. andBlay, D. 1996. New low-Reynolds number k-e model

    including damping effects due to buoyancy in a strati-

    fied flow field. International Journal of Heat and Mass

    Transfer 39(16): 3483-3496.

    Murakami, S. 2004. Analysis and design of micro-climatearound the human body with respiration by CFD.Indoor

    Air 14(Suppl. 7): 144-156.

    Nielsen, P.V., Murakami, S., Kato, S., Topp, C. Yang, J-H.

    2003. Benchmarks test for a computer simulated person.

    Aalborg University, Indoor Environmental Engineering

    (see http://www.cfd-benchmarks.com).

    Pope, S.B. 2000. Turbulent Flows. Cambridge, UK: Cam-

    bridge University Press.

    Posner, J.D., Buchanan, C.R. and Dunn-Rankin, D. 2003.

    Measurement and prediction of indoor air flow in a

    model room.Energy and Building. 35: 515-526.Richardson, L.F. and Gaunt, J.A. 1927. The deferred

    approach to the limit. Philosophy Transactions of the

    Royal Society of London, Ser. A. 226: 299-361.

    Roache, P.J. 1997. Quantification of uncertainty in computa-

    tional fluid dynamics.Annual Review of Fluid Mechan-

    ics 29: 123-160.

    Sideroff, C.N. 2007. Detailed examinations of the human

    micro-environment by CFD.PhD Thesis, Syracuse Uni-

    versity, Dept. of Mechanical and Aerospace Engineer-

    ing.

    Srebric, J., Vokovic, V., He, G. and Yang, X. 2007. CFD

    boundary conditions for contaminant dispersion, heat

    transfer and airflow simulations around human occu-pants in indoor environments. Building and Environ-

    mentArticle in Proof.

    Sveningsson, A. 2003. Analysis of the performance of differ-

    ent -f turbulence models in a stator vane passage flow.

    Ph.D Thesis, Chalmers University of Technology ISSN

    1101-9972.

    Tanabe, S., Kobayashi, K., Nakano, J., Ozeki, Y., and Koni-

    shi, M. 2002. Evaluation of thermal comfort using com-

    bined multi-node thermoregulation (65N) and radiation

    models and computational fluid dynamics (CFD).Energy and Buildings 34: 637-646.

    Tennekes, H. and Lumley, J.L. 1972. A First Course in Tur-

    bulence. Cambridge, MA, USA: MIT Press.

    Topp, C. 2002. Influence of geometry of a computer simu-

    lated person on contaminant distribution and personal

    exposure.ROOMVENT 2002. 1: 265-268.

    Topp, C., Nielsen, P.V., and Srensen, D. 2002. Applicationof computer simulated persons in indoor environmental

    modeling.ASHRAE Transactions 108(2): 1084-1089.

    Xing, H., Hatton, A. and Awbi, H.B. 2001. A study on the air

    quality in the breathing zone in a room with displace-

    ment ventilation. Building and Environment 36: 809-

    820.

    Zhang, T. and Chen, Q. 2007. Novel air distribution systems

    for commercial aircraft cabins. Building and Environ-ment42: 1675-1684.

    Zhu, S., Kato, S., Murakami, S. and Hayashi, T. 2005. Study

    on inhalation region by means of CFD analysis and

    experiment.Building and Environment40: 1329-1336.

    2008, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. (www.ashrae.org).

    Published in ASHRAE Transactions Vol. 114, Part 2. For personal use only. Additional reproduction, distribution, or transmission

    in either print or digital form is not permitted without ASHRAEs prior written permission.