empty1.pdf

download empty1.pdf

of 15

Transcript of empty1.pdf

  • 7/27/2019 empty1.pdf

    1/15

    Building and Environment 43 (2008) 17191733

    The influence of stacks on flow patterns and stratification associated

    with natural ventilation

    Shaun D. Fitzgerald, Andrew W. Woods

    BP Institute for Multiphase Flow, Madingley Rise, Madingley Road, Cambridge CB3 OEZ, UK

    Received 16 September 2007; received in revised form 24 October 2007; accepted 26 October 2007

    Abstract

    We investigate the steady state natural ventilation of an enclosed space in which vent A, located at height hA above the floor, is

    connected to a vertical stack with a termination at height H, while the second vent, B, at height hB above the floor, connects directly to

    the exterior. We first examine the flow regimes which develop with a distributed source of heating at the base of the space. If hBohA, then

    the unique flow solution involves inflow through vent Band outflow through vent A up the stack. IfH4hB4hA, then two different flow

    regimes may develop. Either (i) there is inflow through vent Band outflow through vent A, or (ii) the flow reverses, with inflow down the

    stack into vent A and outflow through vent B. With inflow through vent A, the internal temperature and ventilation rate depend on the

    relative height of the two vents, A and B, while with inflow through vent B, they depend on the height of vent Brelative to the height of

    the termination of the stack H. With a point source of heating, a similar transition occurs, with a unique flow regime when vent Bis lower

    than vent A, and two possible regimes with vent Bhigher than vent A. In general, with a point source of buoyancy, each steady state is

    characterised by a two-layer density stratification. Depending on the relative heights of the two vents, in the case of outflow through vent

    A connected to the stack, the interface between these layers may lie above, at the same level as or below vent A, leading to discharge of

    either pure upper layer, a mixture of upper and lower layer, or pure lower layer fluid. In the case of inflow through vent A connected to

    the stack, the interface always lies below the outflow vent B. Also, in this case, if the inflow vent A lies above the interface, then the lower

    layer becomes of intermediate density between the upper layer and the external fluid, whereas if the interface lies above the inflow vent A,

    then the lower layer is composed purely of external fluid. We develop expressions to predict the transitions between these flow regimes, in

    terms of the heights and areas of the two vents and the stack, and we successfully test these with new laboratory experiments. We

    conclude with a discussion of the implications of our results for real buildings.

    r 2007 Elsevier Ltd. All rights reserved.

    Keywords: Natural ventilation; Stacks; Stratification; Flow regimes

    1. Introduction

    The growing interest in reducing energy demand from

    buildings has stimulated much research in natural ventila-tion. Many of the key principles of natural ventilation have

    been identified using simplified analogue laboratory

    experiments, and supporting theoretical models [15]. The

    majority of these studies focus on buoyancy driven

    displacement ventilation in which relatively cool air enters

    the base of the building, is heated, and then discharges

    from a vent at the top of the building.

    As an approximation, the vent and stack geometry are

    often simplified as being openings at the base and top of

    the building. However, in many real naturally ventilated

    buildings, the vents (e.g. windows) may be located atintermediate levels in a room and there may be substantial

    stack structures which draw air from the building and

    channel this upwards prior to venting to the exterior. In

    buildings with vaulted ceilings and sloping roofs, there may

    be vents or stacks connected to both the lower and the

    upper part of the roof (Fig. 1a). The present study was

    inspired by a new classroom block at the Hagley School in

    Worcestershire in which the top floor classrooms are

    ventilated using a stackvent configuration analogous to

    Fig. 1a. A critical question in the design phase of the

    ARTICLE IN PRESS

    www.elsevier.com/locate/buildenv

    0360-1323/$ - see front matterr 2007 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.buildenv.2007.10.021

    Corresponding author. Tel.: +44 1223 765714.

    E-mail address: [email protected] (S.D. Fitzgerald).

    http://www.elsevier.com/locate/buildenvhttp://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.buildenv.2007.10.021mailto:[email protected]:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.buildenv.2007.10.021http://www.elsevier.com/locate/buildenv
  • 7/27/2019 empty1.pdf

    2/15

    building concerned the height of the stack which would

    optimise the outflow from the classroom. As we identify in

    this work, with the geometry of Fig. 1a, it is possible that

    the stack may involve inflow to the room or outflow from

    the room. We will show that a rich spectrum of natural

    ventilation flows may arise from such a configuration of

    openings, depending on the relative heights of the vents

    and termination of the stack. We also show that with alocalised source of buoyancy, the nature of the temperature

    stratification which develops also depends on the locations

    of the vents and stack.

    In order to develop a systematic understanding of the

    different flow regimes, we explore the natural ventilation in

    a simplified model building with two vents, and in which

    one vent is connected to a stack. We allow the level of both

    vents to vary from the floor to the level of the top of the

    stack (Fig. 1bd). By comparison with Fig. 1a, one can

    recognise that the design of the stack and vents at Hagley

    School, with a sloping roof, is captured by this model

    geometry. In this paper, we explore the flow regimes and

    patterns of internal stratification which develop with such a

    configuration of the vents. To our knowledge, many of the

    flow regimes which we identify have not been described in

    the literature on natural ventilation (eg. [3]).

    Earlier work on the inclusion of intermediate level vents

    [5] has been restricted to the case in which vents are also

    placed at both the top and bottom of a building. In this

    paper we focus on a simplified open-plan type building, in

    which there is a low level heat load and only two vents, one

    at the floor or at an intermediate level, and one which

    accesses a stack venting from the side of the building. The

    case of a distributed heat load, in which the air becomes

    well-mixed within the space, builds on the work of

    Gladstone and Woods [4] who considered the case of a

    room with vents at high and low level in the space. In that

    case, there is a unique upward displacement ventilation

    flow. If in contrast, there are two vents, A and B, at

    intermediate level in the space, with vent A connected to a

    stack which extends to the top of the building, then the

    flow regime depends on the relative height of vents A and

    B. If vent A lies above vent B, then we expect an upwarddisplacement flow in the stack (Fig. 1b). However, if vent B

    lies above vent A, then there may be two different regimes.

    Firstly, the flow may continue to enter the space through

    vent Band exit through vent A rising up the stack (Fig. 1c).

    In this case, the flow is driven by the buoyancy force

    associated with the column of buoyant air in the stack

    above the level of the inflow, vent B. In the second regime,

    the external air enters the stack and descends into the room

    through vent A, while air vents from the space through

    vent B (Fig. 1d). In this case, the flow is driven by the

    buoyancy force associated with the column of air between

    the level of vent A and vent B. The two different flow

    regimes arise from the non-linearity associated with upflow

    and downflow in the stack. In Section 2, we develop a

    model of these multiple states, and, in Section 3, we

    successfully compare our predictions of the different flow

    regimes with some new laboratory experiments. Such

    multiple flow regimes which arise from the non-linearity

    of flow in a stack have been recognised in the different

    context of mixing ventilation through roof mounted stacks

    [6].

    The case in which the room is heated by a localised

    source of buoyancy is more complex and forms the subject

    of the remainder of this paper. In broad-brush terms, as

    with a distributed heat source, there is a transition from a

    ARTICLE IN PRESS

    A

    BhA

    hB

    H

    A A

    BB

    hB

    hBhA

    B

    A

    stack

    window

    Sloping roof

    Internal heat load

    H

    Fig. 1. Schematic of the steady ventilation regimes in a room heated by a distributed source at the base and ventilated by two intermediate level openings,

    one of which is connected to a stack which extends to the top of the room. In (a) we show a schematic of the configuration of the vent and stack at Hagley

    School Worcestershire, in which there is a sloping roof, a high level window on the vertical wall adjacent to the highest point of the roof, and a stack risingabove the vent which connects to the lower side of the roof. In (b)(d) we show a generic building, used for the modelling, which can accommodate the

    stackvent configuration of (a) by suitable choice the vent and stack elevations, but which also allows for the opening of a low level vent/window (Fig. b).

    In (b)(c) the room ventilates in simple upward displacement mode, while in (d), which has the same geometrical configuration as (c), the flow reverses,

    now entering the room through the stack.

    S.D. Fitzgerald, A.W. Woods / Building and Environment 43 (2008) 171917331720

  • 7/27/2019 empty1.pdf

    3/15

    unique upflow displacement regime, in the case that vent

    A connected to the stack lies above the other vent B, to two

    complementary flow regimes when vent A which is

    connected to the stack lies below the other vent B.

    However, with a localised source of heat, a two-layer

    stratification typically becomes established in the space

    [1,2]. In the original work of Linden et al. [2], in whichthere were vents at the top and base of the space, it was

    shown that the interface lies between the inflow and

    outflow vents, that the lower layer is composed purely of

    external fluid and that the outflow fluid is derived purely

    from the upper layer. We show here that with a stack

    connected to one of the vents, there are a number of

    different flow regimes which can develop depending on

    whether the vent connected to the stack lies above or below

    the vent connected directly to the exterior. In each of these

    flow regimes, the interior fluid develops a two-layer

    stratification, but depending on the vent sizes and

    elevations, the outflow may issue from either the lower or

    the upper layer. Similarly, the inflow may enter either the

    upper or lower layer, and in the case in which the fluid

    enters the upper layer, the lower layer becomes of

    intermediate density between the exterior and upper layer.

    In Section 4 we describe a theoretical model which

    categorises these different flow regimes, and we present

    some new analogue laboratory experiments in Section 5 in

    which we demonstrate each of the regimes. We also test our

    predictions, by developing an analogue theoretical model

    for the experiments and comparing the transitions in flow

    regime with that model. In Section 6, we discuss the

    implications of these results for the design of naturally

    ventilated buildings, and consider some avenues for furtherresearch.

    2. Distributed heat loads

    We consider a room in which there is a distributed heat

    load QH at the base of the room which leads to vigorous

    convection and a well-mixed interior [4]. It is assumed that

    there are openings A and B on the sides of the room, of

    area aA and aB, at heights hA and hB above the floor, with

    opening A connected to a stack which rises to a termination

    at elevation Habove the floor, where HXmaxhA; hB. It isimportant to recognise that in this model, H does not

    correspond to the height of the building, but the height of

    the top of the stack. However, with a sloping roof

    configuration, one can imagine that the elevation of vent

    B, on the upper end of the roof, could coincide with the

    elevation of a stack above vent A, located on the lower end

    of the roof.

    We investigate first the flow regime which develops when

    the outflow is through vent A and rises up the stack while

    the inflow is through vent B(Fig. 1b and c). The buoyancy

    driving the flow in this case is associated with a column of

    buoyant room air extending from the level of vent Bto the

    top of the stack, g0H hB, where g0 is the reduced gravity

    of the air in the room, defined as g0 gre rr=re, where

    re and rr are the density of the exterior and interior fluid,

    and g is the acceleration due to gravity.

    If the effective opening area of the two vents is A (cf. [2];

    also see Section 4 herein), then the flow rate V is given by

    V Ag0H hB1=2, (2.1)

    while the heat flux QH is given by the balanceQH rCpDTV, (2.2)

    where DT is the temperature elevation in the room, r is the

    density of air and Cp is the specific heat capacity of air. For

    small changes in temperature,

    g0$gaDT, (2.3)

    where the coefficient of expansion for air a 1=T, T isabsolute temperature expressed in Kelvin, and so the

    temperature elevation of the room is

    D

    T

    Q2H

    ar2C2pA2gH hB !

    1=3

    (2.4)

    and the ventilation rate is given by combining (2.2) and

    (2.4),

    VA2H hBgaQH

    rCp

    1=3. (2.5)

    If the height of vent A, which connects to the stack, lies

    below vent B, then it is possible that the reverse flow regime

    is established in which there is inflow through the stack,

    and outflow through vent B (Fig. 1d). In this case, the

    buoyancy driving the flow is associated with the buoyancy

    of the air in a column between vent A and vent B, given byg0hB hA, and the temperature elevation of the room is

    DTQ2H

    r2C2pA2gahB hA

    !1=3, (2.6)

    while the ventilation is now given by

    VA2hB hAgaQH

    rCp

    1=3. (2.7)

    Fig. 2 illustrates the variation of dimensionless internal

    temperature Dy DTr2C2pA2gaH=Q2H

    1=3 as a function

    of the height of vent B hB hB=H in the case that vent A,which connects to the stack, is located at the points hA 13and 2

    3above the floor of the space, where hA hA=H. The

    two complementary flow regimes, which lead to two

    different temperatures, develop when hA4hB, as expected.

    3. Experimental investigation of multiple flow regimes with

    distributed heating

    Following a similar approach to Gladstone and Woods

    [4], we carried out a series of analogue laboratory

    experiments to test whether the multiple flow regimes

    predicted in Section 2 do indeed develop in an analogue

    experimental system. Using water as the working fluid, we

    ARTICLE IN PRESS

    S.D. Fitzgerald, A.W. Woods / Building and Environment 43 (2008) 17191733 1721

  • 7/27/2019 empty1.pdf

    4/15

    immersed a perspex tank of height 28.6cm and area

    17:8 cm 17:8 cm, with walls of thickness 0.8 cm, in a largereservoir of water. The experimental tank included a

    number of openings at different levels on the side of the

    tank, and there was also a central vent, which connectedthe room to the exterior through a stack. To model a

    distributed source of heating, we used a coiled high

    resistance wire placed just above the floor of the tank.

    The heat flux produced could then be calculated from

    measurement of the current through and voltage loss

    across the wire [6]. Note that this is somewhat different to

    the technique used by Gladstone and Woods [4] in which

    the base plate was maintained at a constant temperature

    and the associated heat flux was inferred from empirical

    laws for turbulent thermal convection. The temperature

    distribution in the tank was recorded by type K thermo-

    couples, of accuracy 0:1 C, which were spaced at regularvertical intervals of 3 cm within the tank. With sufficient

    heat load, in the range 200500 W, this experimental

    system leads to turbulent thermal convection in the

    experimental room, with Rayleigh numbers of order

    108109. At these values of the Rayleigh number the fluid

    within the tank is well-mixed and isothermal, as measured

    directly from thermocouples in the tank [4], thereby

    providing a good analogue to the well-mixed ventilation

    regime within a room heated by a distributed source of

    buoyancy. The flow through the openings, of diameter

    12 cm, with speed 510 cm/s, has Reynolds number of

    order 10002000. In order to apply the model of Section 2

    to such flows, we require an estimate of the loss coefficient

    for the openings. The effective combined loss coefficients

    for the vents and stacks were measured using a calibration

    experiment in which the tank was filled with a saline

    solution and then immersed in the reservoir of fresh water.

    Rubber plugs were removed from two of the ventilation

    openings, at high and low level, and the rate at which the

    solution then drained from the tank was recorded (cf. [4]).As described in the Appendix, the effective loss coefficient

    in the calibration experiment was found to be $0:6, withtypical Reynolds numbers in the stack of order 1000. We

    therefore use this value in comparing our experimental

    data with the predictions of the model of Section 2. In

    principle, one can model the frictional losses through both

    the pipe, which serves as the stack, and the opening from

    the tank into the pipe [7], but for the range of Reynolds

    numbers in the present experiments, the empirical value for

    the loss coefficient of 0:6 is sufficiently accurate. By makingthese measurements, and demonstrating that the constant

    value for the loss coefficient describes this experimental

    data, we have established that our model of the flow

    through the experimental stack system is analogous to that

    for a real building; in applying the model to a real building

    one would however need to determine the loss coefficient

    which pertains to the actual building. Note that in this

    experimental system, the heat losses from the walls of the

    experimental tank are small compared to the heat flux

    convected with the ventilation flow (cf. [4,7]). We infer that

    the flow regimes which develop within the experimental

    model are dynamically analogous to those in a real

    building (cf. [3,4]).

    We then conducted a systematic series of experiments to

    explore the steady state flow regimes as a function of theelevation of vent B, with a fixed height of vent A which

    connects to the stack of 12cm. Experiments were

    conducted for a range of heat fluxes, and both the inflow

    and outflow regimes in the stack were realised. In Fig. 2, we

    compare the experimental measurements of the dimension-

    less temperature excess in the space,

    Dy DTA2r2C2pgHaw

    Q2H

    !1=3, (3.1)

    where aw is the coefficient of expansion for water, as a

    function of the dimensionless height of the vent B, hB.

    According to the model of Section 2,

    Dy 1

    1 hB1=3

    , (3.2)

    for the stack outflow mode, while

    Dy 1

    hB hA1=3

    , (3.3)

    for the stack inflow mode. It is seen in Fig. 2 that there is

    very good agreement between the predictions and the

    experimental results. Note, however, that as the vertical

    separation between the mid-points of the two vents

    becomes small relative to the vertical extent of the

    ARTICLE IN PRESS

    0.5

    1

    1.5

    2

    2.5

    0 0.2 0.4 0.6 0.8 1

    hB

    Fig. 2. Non-dimensional temperature Dy as a function of the non-

    dimensional height of vent B. The thick solid line corresponds to outflow

    from vent A and through the stack and the thin solid and dotted lines to

    inflow through the stack for hA 13

    and 23, respectively. Also, shown is

    comparison of experimental data (symbols) with theory for the case

    hA 0:58. The dot-dashed line and open squares correspond to inflow andblack squares to outflow through vent A.

    S.D. Fitzgerald, A.W. Woods / Building and Environment 43 (2008) 171917331722

  • 7/27/2019 empty1.pdf

    5/15

    openings, the nature of the flow through each vent changes

    from uni-directional to bi-directional and the model is then

    no longer valid.

    4. Localised source of buoyancy

    We now turn to the more complex case in which the heatsource is localised rather than distributed over the floor of

    the space. We explore the effect of the height of the two

    vents, and the presence of a stack, on the ventilation

    regime. A localised source of heating at the base of the

    room typically generates a turbulent buoyant plume and if

    the space is ventilated through vents at the top and base of

    the room, then the interaction of the plume with the

    ventilation leads to the formation of a two-layer stratifica-

    tion within the space [2]. In this flow regime, the interface

    lies between the two vents, the lower layer is typically

    composed of pure external fluid and only upper layer fluid

    vents from the space. We now show how variations in the

    height of the two vents, and the presence of a stack

    connected to one of the vents, can change the internal

    stratification and flow regime substantially.

    For our analysis, we assume there is a stack connected to

    a vent A, located at an intermediate height, hA, while the

    other vent, B, connected directly to the exterior, has height

    hB (cf. Fig. 1). We assume there is a localised heat source of

    magnitude QH at the base of the room. We explore the

    different flow regimes which may develop as the elevation

    of vent B rises from the base of the room to the top of the

    room, and we focus on the case hBoH, where H is the

    height of the termination of the stack, as is typically the

    case in practice. Again, we emphasise that the height of the

    top of the stack H does not need to correspond to theheight of the top of the room; indeed, even in the case that

    vent Bhas the same elevation as the top of the stack above

    vent A, one can imagine a building with a sloping roof,

    with the stack connected to vent A on the lower part of the

    roof, and a simple vent B connected to the upper part of

    the roof (cf. Fig. 1a).

    The results for the case A 0:1 are shown in the regimediagram of Fig. 3(a), where A A%=l3=2H2 with l % 0:12for a fully developed turbulent buoyant plume [8], A%

    cAaAcBaB=12c2Aa

    2A c

    2Ba

    2B

    1=2 and where c is the loss

    coefficient for the vent. When vent B is located at the base

    of the room, two regimes, which we denote as I and V, may

    develop depending on the elevation of vent A above the

    base of the room relative to the theoretical height of the

    fluid interface, hL, as predicted by the theory of Linden

    et al. [2]. If vent A is located sufficiently high in the room,

    then the interface is predicted to lie below vent A, hLohA,

    and there is pure outflow of the upper layer fluid, while

    the lower layer is composed of external fluid (regime I,

    ARTICLE IN PRESS

    I

    V

    II

    IV

    III

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 0.2 0.4 0.6 0.8 1

    hB

    hA

    A

    BB

    A

    A=0.1

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 0.2 0.4 0.6 0.8 1

    II

    IV

    V

    III

    I

    hB

    hA

    A=0.001

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 0.2 0.4 0.6 0.8 1

    III

    IV

    V

    I II

    hB

    hA

    A=1

    Fig. 3. Regime diagram illustrating the different flow regimes which develop in the case of a localised heat source for the cases in which (a) A 0:1,

    (b) A 0:001, (c) A 1. In each case, the stack provides the conduit for the outflow.

    S.D. Fitzgerald, A.W. Woods / Building and Environment 43 (2008) 17191733 1723

  • 7/27/2019 empty1.pdf

    6/15

  • 7/27/2019 empty1.pdf

    7/15

    state, this is not possible since there would be no loss of

    buoyant fluid from the room. Instead, the interface depth

    becomes fixed at hA in order that buoyant fluid can vent

    from the space. The density of the upper layer then equals

    that of the plume as it reaches the lower interface of this

    layer. However, now the volume flux supplied to the upper

    layer by the plume is smaller than the buoyancy drivenoutflow through the stack associated with fluid of that

    buoyancy. As a result, to maintain a steady state, lower

    layer fluid is also convected up the stack. This decreases the

    buoyancy of the fluid in the stack. This regime is denoted

    as V in the regime diagram (Fig. 3a).

    In this equilibrium regime, the volume flux in the stack V

    is given by the sum of the flux in the plume Vp at height hA,

    Vp lb1=3h

    5=3A , (4.5)

    and the flux from the lower layer, Vl into the stack, where b

    is the buoyancy flux at the base and l 0:12 as before. The

    buoyancy of the fluid in the stack g

    0

    s, which is assumed tobe homogenous, is given by the buoyancy of the plume at

    height hA, g0p b=Vp, multiplied by the factor Vp=Vp

    Vl owing to the dilution by the lower layer (ambient) fluid

    which enters the stack. By matching the volume flux

    associated with the outflow of fluid from the stack

    A%ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    g0sH hAp

    with the sum of the volume flux supplied

    to the upper layer by the plume and Vl, we deduce that the

    ratio of the flux from the lower layer to the total fluid

    entering the stack, Vl=Vl Vp, is given by the relation,

    Vl

    Vl Vp 1

    h5=3A

    A2=31 hA1=3

    . (4.6)

    The variation of Vl=Vl Vp with hA is shown in Fig. 4for several values of A. As the ratio Vl=Vl Vpincreases, the buoyancy of the outflow decreases, since

    the fluid is being diluted with progressively more lower

    layer ambient fluid. As expected, it is seen that Vl=Vl Vp increases with the effective opening area, A, and with

    decreasing height of the point of access to the stack hA.

    Note that in practice, when there is outflow from both

    layers, the interface elevation far from the outflow vent

    may be a little different from that at the vent. This is

    because near the outflow the fluid is moving much more

    rapidly than far from the vent and this can lead to a

    reduction in the fluid pressure. Indeed, in the experimentsreported in the next section, in which H 20230cm, we

    find that in this mixed flow regime there is a small

    difference, o0:5 cm, between the far-field height of theinterface and the elevation of vent A, hA.

    4.2.2. Transition VIII

    We now consider how the internal stratification evolves

    as the height of the inflow vent Brises from the base of the

    room in the situation in which the flow regime commences

    in regime V. The discussion corresponds to moving along

    line BB0 in Fig. 3a. As hB rises, the interface remains fixed

    to the level of the outflow vent, hA. Eventually, hB reaches

    this same level and then rises above the interface and

    outflow vent A, so that the flow adjusts to regime III as

    described above and shown in Fig. 3a.

    Now, the incoming fluid will enter the upper layer and,

    being relatively dense, will descend through this layer. As itdescends, it will entrain some of the upper layer fluid before

    reaching the interface. As a result, it supplies the lower

    layer with fluid of intermediate density. While the interface

    remains fixed at the level of the outflow vent A, the fluid

    which vents from the room is composed of a mixture of

    upper and lower layer fluid. As the height of the inflow vent

    B increases, the lower layer becomes progressively more

    buoyant, and the fraction of lower layer fluid venting from

    the space increases. Eventually, a critical point is reached at

    which all the fluid which vents from vent A and through the

    stack originates from the lower layer. IfhB, the elevation of

    the inflow vent B increases any further then this will cause

    the interface to rise above the level hA corresponding to the

    height of the outflow vent A. At this critical point, the flow

    regime evolves to regime IV, as shown in Fig. 3a.

    4.2.3. Regime transition IIIIV

    We now develop a simple model to predict the point of

    transition between regime III and IV, which occurs when

    the interface is located at the level of the outflow stack, hA,

    but all the outflow is derived from the lower layer. To

    model this transition, we note that the conservation of

    buoyancy in the room may be written as

    b Vg0l (4.7)

    ARTICLE IN PRESS

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 0.2 0.4 0.6 0.8 1

    hA

    Vl/(Vl+Vp)

    Fig. 4. Variation of the proportion of lower layer fluid which vents from

    the stack as a function of the height of access to the stack hA when a room

    is heated by a point source at the base and ventilated by an opening in the

    base and an opening on the side wall which connects into a stack. Curves

    are given for dimensionless vent areas A 0:1, 0.5, 1 and 2 by the solid,dashed, dotted and dot-dashed lines, respectively.

    S.D. Fitzgerald, A.W. Woods / Building and Environment 43 (2008) 17191733 1725

  • 7/27/2019 empty1.pdf

    8/15

  • 7/27/2019 empty1.pdf

    9/15

    upper layer fluid is given by

    hA

    1

    2 hB A

    2=5

    1 hB

    1=5

    . (4.21)The solution of this relation is also shown on the regime

    diagram (Fig. 3a). Again, we find that in the limit hB ! 1

    then Eq. (4.21) requires that hA again decreases, owing to

    the non-linear parameterisation of the mixing in the fluid

    which flows in through the inflow opening. Indeed, in the

    limit hB ! 1 we find that hA ! 0:5. Again, although theoverall predictions are consistent with our experiments, we

    have not been able to test this detailed prediction using our

    experimental system owing to limitations of the size of our

    apparatus (Section 5).

    The regime diagram shown in Fig. 3a corresponds to the

    case A 0:1. In Figs. 3b and c we show the regime diagramfor the cases in which A 0:001 and 1, respectively. It canbe seen that in the case A51, for which the room is

    effectively very tall relative to the typical dimension of the

    openings, then regimes II and IV dominate, whereas when

    A is larger, the other, intermediate regimes can be observed

    for a wider range of values of hA and hB. Note that in the

    case of very small A, the assumption that there is no heat

    loss through the fabric of the ventilated space becomes

    less accurate, and, in the case of outflow through the stack

    (Fig. 3b) this will lead to a reduction in the temperature of

    the upper layer and hence the buoyancy force and overall

    ventilation flow rate. Modelling such effects in detail is

    beyond the scope of the present study, but some of theeffects of such heat losses have recently been described by

    Livermore and Woods [13].

    4.3. Reverse flow regime with inflow through the stack and

    vent A and outflow through vent B

    As mentioned in the introduction, once the elevation of

    vent A, connected to the stack, lies below that of vent B

    which is connected directly to the exterior, it is possible for

    the flow to reverse in direction, with inflow through the

    stack and vent A and outflow through vent B. In this flow

    regime, the interface either lies above (regime i, Fig. 5) or

    below (regime ii, Fig. 5) the inflow vent A at the base of the

    stack, where the cold ambient fluid enters the room. In

    regime i, the ventilation of the room is similar to the classiccase described by Linden et al. [2], but with a modified

    height hB hL over which the buoyancy acts to drive the

    flow, where hL is the height of the interface within the

    room. Consequently, hL is given by a modified form of

    (4.1)

    A h5L

    hB hL

    1=2. (4.22)

    In order that the interface lies above the inflow vent, we

    require hAohL where hL is given by (4.22). If the interface

    lies below the point of access to the stack then regime ii

    develops whereby the incoming fluid will enter therelatively hot upper layer and descend. As the relatively

    dense ambient air descends, it will form a plume and

    entrain some of the hot upper layer. Consequently, the

    lower layer will be warmer than the exterior fluid. The

    delineation between this regime ii and regime i is shown in

    the regime diagram (Fig. 6), with the transition from

    regime i to ii being given by setting hA hL in Eq. (4.22).

    5. Experimental observations

    We have developed a series of new analogue laboratory

    experiments to investigate whether the different flow

    regimes that we described in Section 4 actually develop in

    practice. In the experiments of Section 3, we modelled a

    distributed source of heating using a heated wire. However,

    in order to model the convective flow associated with a

    localised heat source, it is easier to use the analogue system

    of fresh and saline water as described by Baines and Turner

    [10]. The focus of the experiments is to demonstrate that

    the multiplicity of different flow regimes predicted in

    Section 4 can indeed develop in an analogue experimental

    system. We also develop the model of Section 4 for

    application to the experiments, now using properties of a

    saline plume in water, and we compare the experimental

    observations of the conditions for transitions in the flow

    ARTICLE IN PRESS

    REGIME i

    hA

    REGIME ii

    hBhA

    Fig. 5. Schematic diagram showing the two layer stratification in the room when inflow through the stack occurs. Regime i occurs when the interface in

    the room lies above the base of the stack and regime ii when the interface lies below the base of the stack.

    S.D. Fitzgerald, A.W. Woods / Building and Environment 43 (2008) 17191733 1727

  • 7/27/2019 empty1.pdf

    10/15

    regime with the predictions of this model. Similar experi-

    mental models exploring the dynamics of saline turbulent

    plumes within a ventilated, enclosed chamber have been

    applied as an analogue model for natural ventilation flows

    from localised sources of heating in a number of earlier

    studies (cf. [3,11]). The turbulent buoyant plumes gener-

    ated from the localised source of saline water within themodel room represent an analogue for the natural

    convective flows generated from localised heat sources

    within a building, provided that the source mass flux is

    much smaller than the ventilation flow, and that the plume

    is of sufficient Reynolds number to be fully turbulent [11].

    The experimental system consisted of a small perspex

    tank (the room) of height 28.6 c m and area

    17:8 cm 17:8 cm. This was partially immersed in a largereservoir of height 48 cm and area 88cm 43 cm, which

    acted as the exterior environment. We used fresh water in

    the environment and saline solution as the source of

    (negatively) buoyant fluid. The system was therefore run

    upside down, with the plume source located below the free

    surface of the water in the tank, and the stack termination

    being located above the base of the tank (Fig. 7). The dense

    saline solution injected through the plume source, which

    moves downwards, corresponds to a source of hot air risingfrom a localised heat source on the base of a real room.

    One side wall of the experimental tank had a large

    number of circular ventilation holes, of diameter 5, 6 and

    15 mm. These ventilation holes were sealed by rubber

    stoppers. These could be removed to model a wide range of

    effective vent areas and heights between the model room

    and the exterior reservoir. A hole in the base of the tank

    also enabled stacks of internal diameter d, and with the

    entry point being located at a height below the roof of the

    tank, ha, taking values d; hA given by (13 mm, 189 mm),(13 mm, 245 mm), (10 mm, 208 mm), (10 mm, 230 mm). The

    stack entrance for the fluid was the horizontal plane of the

    top of the cylindrical pipe used for the stack.

    The room and exterior tank were initially filled with

    fresh water. Then, for convenience a 5% salt solution was

    supplied from above at a constant rate 0.3 cc/s via a twin-

    feed peristaltic pump, as the source for a descending

    (negatively) buoyant plume. The nozzle used for the plume

    source was described by Woods et al. [11]. The room was

    elevated 20 cm from the base of the exterior tank so that

    the dense fluid exiting from the room could sink to the base

    of the reservoir, away from the room. A systematic series

    of experiments were conducted with the plume source 7.1,

    10.8, 13 and 14.5 cm from the base of the tank, and with the

    number of upper vents ranging from one 6 mm diametervent to, at most, 4 mm 6 mm and 3 mm 15mm

    diameter vents all being open simultaneously. As in Section

    3, the effective combined loss coefficients for the vents and

    stacks were found to be $0:6 with typical Reynolds

    ARTICLE IN PRESS

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 0.2 0.4 0.6 0.8 1

    iiii

    i

    hB

    hA

    Fig. 6. Regime diagram illustrating the different flow regimes which

    develop, in the case of a localised heat source, for the case in which

    A 0:1 and when the stack is the source of inflow.

    Reservoir

    tank

    Experimental

    model room

    Source of saline

    fluid to pump

    into tank

    Peristaltic Pump

    Plume

    source

    Pipe connected

    to the opening in

    base represents

    analogue of a stack

    Ventilation

    opening in

    side of model

    room

    Free surface

    Supports to

    raise model

    room off floor

    of reservoir

    Fig. 7. Schematic of the experimental apparatus, illustrating how the model room is immersed within the external reservoir tank, and the geometry of the

    inflow nozzle and the experimental stack.

    S.D. Fitzgerald, A.W. Woods / Building and Environment 43 (2008) 171917331728

  • 7/27/2019 empty1.pdf

    11/15

    numbers of order 1000. This is discussed further in the

    Appendix. In comparing the experimental observations of

    the transition in flow regime with the predictions of the

    theoretical model, we adopt the model of Section 4, but

    now use physical properties of water and aqueous saline

    solution. The main comparison of the experimental results

    with the theoretical predictions is in testing the flow

    regimes which develop with different plume source heights

    and with different net sizes of ventilation opening;

    however, we also compare the interface height observed

    in the experiments with the theoretical predictions (Fig. 8a)

    and the density of the outflowing fluid with the theoretical

    predictions (Fig. 8b).

    5.1. Regimes I and V

    In the first series of experiments a stack was located in

    the base of the tank whilst the number of ventilation holes

    at the top of the tank was systematically increased. For

    each configuration, we fixed (a) the distance between the

    plume source nozzle and external termination of the stack,

    H, which in our experiments also corresponds to the depth

    of the room, and (b) the stack height hA, and we

    determined the critical value of A% at which there was a

    transition in the flow regime, as may be seen in themeasurements of the steady state height of the interface

    (Fig. 8a). The transition point corresponds to a change in

    regime from single layer outflow (regime I, Fig. 3) to mixed

    outflow with the interface fixed at the level of vent A

    connected to the stack (regime V, Fig. 3). The photographs

    in Fig. 9 illustrate the two different regimes, with Fig. 9a

    showing the case with both upper and lower layer fluid

    supplying the stack and Fig. 9b the case of only fluid from

    the buoyant layer in the stack. Note that in Fig. 8a we have

    combined the experimental results for a range of different

    stack heights by plotting hL as a function ofA. This allows

    the theoretical prediction of interface height for the single

    layer outflow mode (regime I, Eq. (4.1)) to be compared

    with experimental observations whilst clearly separating

    the transition points for different stack heights.

    The proportion of upper and lower layer fluid in the

    outflow from the stack can be deduced from measurements

    of salt concentration in the stack and the lower saline layer.

    In Fig. 8b we show how this varies as a function of effective

    vent area in the mixing outflow regime. Note that in each

    case, for small vent area the outflow is derived purely from

    the lower layer of saline fluid and the interface lies above

    the outflow stack in the experiment (regime I, Fig. 3).

    However, with larger area, the interface is fixed at the level

    of the access point to the outflow stack and the outflow is amixture of fluid from the upper ambient layer and lower

    saline layer.

    To compare the flow behaviour directly with the

    predictions of our model we need to account for the effect

    of a virtual origin since the plume source has finite mass

    flux (cf. [11,12]). The correction method suggested by Hunt

    and Kaye [12] has been used, and it may be seen in Figs. 8a

    and b that the model predictions are in good accord with

    the experimental observations. In particular the agreement

    of interface height in the cases for which the interface lies

    below the level of the outflow vent, as shown in Fig. 8a,

    suggests that the experimental plumes are well modelled by

    the theory of turbulent plumes (cf. [2]).

    5.2. Intermediate level vent and intermediate access point to

    stack

    In the main series of experiments, the plume source was

    located just below the free surface of the water in the room,

    1317 cm above the floor of the room (Fig. 7), whilst a

    series of stacks of different heights and a series of vents

    connecting directly to the exterior were used in order to

    map out all the different flow regimes predicted in Fig. 3.

    Five percent of saline solution was injected from the plume

    source at a rate of 0.3 cc/s and each experiment allowed to

    ARTICLE IN PRESS

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 0.2 0.4 0.6 0.8 1

    Vl/(Vl

    +Vp

    )

    hA=0.16

    hA=0.16

    hA=0.36

    hA=0.3

    hA=0.48

    hA=0.22

    hA=0.36

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0 0.1 0.2 0.3

    hL

    A

    A

    Fig. 8. (a) Variation of interface height hL as a function of vent area A for

    the case in which the room is ventilated by a stack and a vent at the height

    of the plume source. The thick solid grey line corresponds to Eq. (4.1) cf.

    Linden et al. [2]. The other lines indicate the stack heights hA used in the

    experiments, results of which are represented by symbols. Open squares

    correspond to hA 0:16, filled squares to hA 0:22, open triangles tohA 0:3, filled triangles to hA 0:36 open diamonds to hA 0:48 andfilled diamonds to hA 0:57. (b) Variation of proportion of ambient fluidin the stack as a function of A for mixed outflow mode. The solid and

    dashed lines denote the predictions from (4.6) for hA 0:16 and 0.36,respectively, and the open squares and filled triangles represent the

    corresponding experimental results. Experimental errors due to measure-

    ment of interface height are estimated to be 0.5 cm, which is at most 5%.

    S.D. Fitzgerald, A.W. Woods / Building and Environment 43 (2008) 17191733 1729

  • 7/27/2019 empty1.pdf

    12/15

  • 7/27/2019 empty1.pdf

    13/15

    this case, the interface always lies below the outflow vent.

    However, depending on the location of the inflow vent

    supplied by the stack, the interface may lie above or below

    the inflow vent, and therefore the lower layer is composed

    of either pure external fluid or fluid of intermediate

    composition.

    ARTICLE IN PRESS

    Fig. 10. Photographs from seven experiments in which the room was ventilated by a stack accessed at an intermediate level and an intermediate level vent.

    Cases shown for internal stack diameter 13.5 mm, intermediate level vent 15mm diameter. (a) Regime Ionly upper layer fluid vents from the stack, cold

    lower layer (hA 0:46, hB 0:15); (b) Regime Vfluid from both the lower and upper layers vents from the stack, cold lower layer (hA 0:25, hB 0:1);(c) Regime IIonly upper layer fluid vents from the stack, warm lower layer (hA 0:46, hB 0:77); (d) Regime IVonly lower layer fluid vents from thestack, warm lower layer (hA 0:16, hB 0:5); (e) Regime IIIfluid from both the lower and upper layers vents from the stack, warm lower layerhA 0:33; hB 0:38; (f) Regime iinflow through stack, cold lower layer (hA 0:21; hB 0:94); (g) Regime iiinflow through stack, warm lower layer(hA 0:46; hB 0:77). All experiments are conducted upside down c.f. a heated room and in the descriptions above cold refers to fresh water in theactual experiment and warm to saline water. Dotted lines indicate the position of the interfaces, arrows the direction of flow through the vents.

    S.D. Fitzgerald, A.W. Woods / Building and Environment 43 (2008) 17191733 1731

  • 7/27/2019 empty1.pdf

    14/15

    The flow regimes predicted by the models have been

    demonstrated in a series of new analogue laboratory

    experiments, for both the distributed and the point sources

    of buoyancy. Furthermore, we have developed the

    theoretical models for application to the laboratory

    experiments, and found reasonable agreement between

    the experimental observations of transitions betweenregimes and the prediction of the model.

    The work is significant in that it exposes the dominant

    control that the elevation of the ventilation openings, and

    the presence of stacks, may have on natural displacement

    ventilation flows. The situation is of relevance, since many

    naturally ventilated buildings are designed to have outflow

    through stacks. However, we have shown that a range of

    different flow patterns can develop with the stacks

    providing a pathway for either inflow or outflow, and that

    in some situations, both of these different flow regimes may

    develop. As well as stacks and vents designed for natural

    ventilation, the work has implications for the natural flows

    which may develop when a room exchanges air through a

    combination of a chimney and window.

    In some buildings with flat roofs, stacks extend above the

    height of the building, to height H above the base of the

    room. In this case, the dimensionless height of the vent (e.g.

    a window) which opens directly to the environment, scaled

    relative to H, hB is always smaller than unity. As a result,

    only a sub-set of the flow regimes described herein ( Figs. 3

    and 6) are accessible. However, in other buildings, for

    example, those with sloping roofs, as at the Hagley School

    in Worcestershire, the height of the opening which connects

    directly to the exterior may be intermediate to the height of

    the opening to the stack from the interior space and the

    height of the top of the stack. In this case, all the regimes

    may be accessible.

    In addition to the importance of recognising these

    regimes for developing control systems for the flow, and

    in particular for locating sensors to detect the different flow

    regimes, these findings may also be of importance formodelling the dispersal of smoke or chemicals from a point

    source release within a building, and for predicting the

    concentration of contaminants issuing from the building.

    Appendix

    The effective loss coefficient for a vent connected to a

    stack was determined experimentally. The tank with an

    open top and a stack connected to the base was filled with

    5% saline solution. The external reservoir was then

    carefully filled such that the water level was 7 cm above

    the top of the experimental tank. A loosely placed rubberbung which was used to seal the stack whilst the apparatus

    was filled was then dislodged and the descending interface

    recorded by video using the shadowgraph technique. The

    effective area A was determined by comparing the data

    with the theoretical prediction [3] and fitting the loss

    coefficient. The best fit was obtained for a loss coefficient of

    0.6 as shown in Fig. A1. This example corresponded to a

    stack of vertical extent 9.5 cm and diameter 13 mm.

    The deviation from the theoretical prediction at late time

    time 130s corresponded to the time at which the

    interface approached the top of the stack. At this time

    the fluid exiting from the stack became a mixture of fresh

    and saline fluid and the draining box model [3] was

    therefore no longer appropriate to describe the ventilation

    flow.

    References

    [1] Sandberg M, Lindstrom S. Stratified flow in ventilated roomsa

    model study. In: Proceedings of roomvent 2nd international

    conference on air distribution in rooms, Oslo, Norway; 1990.

    [2] Linden PF, Lane-Serff GF, Smeed DA. Emptying filling boxes: the

    fluid mechanics of natural ventilation. Journal of Fluid Mechanics

    1990;212:30935.

    [3] Linden PF. The fluid mechanics of natural ventilation. Annual

    Review of Fluid Mechanics 1999;31:20138.

    ARTICLE IN PRESS

    0

    0.2

    0.4

    0.6

    0.8

    0 0.2 0.4 0.6 0.8 1

    I

    V

    II

    IV, i

    II, ii

    IV, ii

    III, i III, iihB

    hA

    Fig. 11. Regime diagram for the laboratory experiments in which the

    room is ventilated by a stack accessed at an intermediate level and anintermediate level vent. The case shown is for A 0:13. Experimentalobservations of the flow regime are indicated by symbols: Diamonds

    correspond to regime I, open circles to V, filled squares to II, open squares

    to IV, filled circles to III, triangles to regime i and crosses to ii.

    0

    5

    10

    15

    20

    25

    30

    35

    0 50 100 150 200 250 300 350

    time (s)

    distance(cm)

    Fig. A1. Experimental measurements of height of interface as a function

    of time (diamonds) compared with the theoretical prediction (solid line)

    for the case in which the loss coefficient for the stack and vent 0:6.

    S.D. Fitzgerald, A.W. Woods / Building and Environment 43 (2008) 171917331732

  • 7/27/2019 empty1.pdf

    15/15

    [4] Gladstone C, Woods AW. On buoyancy-driven natural ventilation of

    a room with a heated floor. Journal of Fluid Mechanics

    2001;441:293314.

    [5] Fitzgerald SD, Woods AW. Natural ventilation of a room with vents

    at multiple levels. Building and Environment 2004;39:50521.

    [6] Chenvidyakarn T, Woods A. Multiple steady states in stack

    ventilation. Building and Environment 2005;40:399410.

    [7] Livermore S, Woods AW. Natural ventilation of a building withheating at multiple levels. Building and Environment 2007;42:141730.

    [8] Morton BR, Taylor GI, Turner JS. Turbulent gravitational convec-

    tion from maintained and instantaneous sources. Proceedings of

    Royal Society of London Series A 1956;234:123.

    [9] Turner JS. Buoyancy effects in fluids. Cambridge: Cambridge

    University Press; 1979.

    [10] Baines WD, Turner JS. Turbulent buoyant convection from a

    source in a confined region. Journal of Fluid Mechanics 1969;37:

    5180.

    [11] Woods AW, Caulfield CP, Phillips JC. Blocked natural ventilation:

    the effect of a source mass flux. Journal of Fluid Mechanics

    2003;495:11933.[12] Hunt GR, Kaye NG. Virtual origin correction for lazy turbulent

    plumes. Journal of Fluid Mechanics 2001;435:37796.

    [13] Livermore S, Woods AW. On the effect of distributed cooling in

    natural ventilation. Journal of Fluid Mechanics 2007, in press.

    ARTICLE IN PRESS

    S.D. Fitzgerald, A.W. Woods / Building and Environment 43 (2008) 17191733 1733