1-s2.0-S0307904X06002927-main

download 1-s2.0-S0307904X06002927-main

of 13

Transcript of 1-s2.0-S0307904X06002927-main

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    1/29

    Mathematical modeling of molten metal dispensing: A studyof a pneumatically actuated diaphragm-driven pump

    Vivek Jairazbhoy   a, Randy C. Stevenson   b,*

    a Climate Control, Automotive Components Holdings, Plymouth, MI 48170, USAb Chassis CAE, Automotive Components Holdings, Dearborn, MI 48120, USA

    Received 1 June 2005; received in revised form 1 August 2006; accepted 2 November 2006Available online 19 January 2007

    Abstract

    This study describes a lumped-element model for a pneumatically actuated, diaphragm-driven pump. The pump wasdeveloped to dispense molten metal for automotive electronic circuit manufacturing. In place of a CFD analysis, whichproved to involve excessively long computations because of a large range of length scales, we created a lumped-elementrepresentation of the physics. The pump has five main region volumes: an air filling chamber, a solder reservoir, an adjust-able throttle valve, a main pumping chamber, and a dispensing nozzle or set of nozzles. The pumping process employs acompressed gas pulse, which fills the air chamber and deforms a diaphragm, which then displaces molten solder in the mainpumping chamber. The description of the dispensing process requires the coupling of three physical domains: pneumatic,

    mechanical, and hydraulic. The fluid motion in the pump is modeled by creating lumped fluid elements from control vol-umes, some with deformable surfaces. The equations for the fluid elements are derived from Bernoulli’s equation, whenviscous effects are not important, or from the general integral momentum balance, when viscous effects are important.A parametric study describing the relationships between variables is presented, and experimental data is used to confirmthe validity of the model.  2006 Published by Elsevier Inc.

    1. Introduction

    We present a mathematical model of the operation of a pump designed for use as a molten solder dispenser

    in electronic manufacturing applications. Fig. 1a shows a schematic diagram of the pump, an early conceptdrawing, and Fig. 1b shows a picture of a later actual working prototype. The pump operates by displacingmolten metal with a pneumatically actuated diaphragm. The resulting dispense characteristics, such as the dis-pense volume, and the dispense timing, are sensitive to numerous parameters, geometrical and operational. Arational selection of these parameters can be significantly aided, if the pump operation can be modeled.

    0307-904X/$ - see front matter    2006 Published by Elsevier Inc.

    doi:10.1016/j.apm.2006.11.005

    * Corresponding author.E-mail addresses: [email protected] (V. Jairazbhoy), [email protected] (R.C. Stevenson).

    Applied Mathematical Modelling 32 (2008) 141–169

    www.elsevier.com/locate/apm

    mailto:[email protected]:[email protected]:[email protected]:[email protected]

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    2/29

    Nomenclature

    asonic   characteristic (inverse time) for sonic flowthrough solenoid

    asubsonic  characteristic (inverse time) for subsonicflow through solenoid

    c p   heat capacity of filling gas at constantpressure

    C S   office discharge coefficient of solenoidDP viscous pressure drop at a port due to viscous lossd D   diameter of diaphragmoV    boundary of fluid deviceoV ports   boundary of all ports of fluid deviceoV walls   boundary of wall of fluid deviceE    elastic modulus for diaphragm materialg   an adjustable constant used to gauge the

     jump condition in flow rates before andafter the diaphragm hit the stop

    f    body force field per unit massF  p   face of port p  of fluid deviceF wall   face of wall of fluid deviceu   fluid velocity potentialU p   average velocity potential over port p g    acceleration due to gravityG N,  G T   frictional force (due to viscous losses) in

    the nozzle and throttle volume regions,respectively

    ret

    effective ‘‘G ’’ for retraction phaseGwall   force of fluid on wall through viscous andpressure forces

    G R   surface tension force in the nozzle regionh pq   vertical height between port p and port  qhN   vertical height of nozzlehR   vertical height of reservoirhT   vertical height of throttlek D   the effective spring constant for the dia-

    phragmk in,   k out adjustable dimensionless constants char-

    acterizing the degree of loss of kinetic en-

    ergy at a port due to viscous surface forcesK  p   kinetic energy factor for port p K drvN   ; K 

    drvT   effective ‘‘K ’’ for the driving phase for

    the nozzle stream and throttle streamrespectively

    K ret effective ‘‘K ’’ for retraction phasel  pq   characteristic length scale along flow from

    port p  to port  ql N   flow path length of nozzlel R   flow path length of reservoirl T   flow path length of throttle

    l̂    unit vector along direction of centralstream line in lumped fluid element

    M   matrix of connection coefficientsM  pq   connection coefficient between port p  and

    port  q M eff  pq   effective connection coefficient from port p

    to port  ql   coefficient of viscosity for fluidn p   normal vector to face  F  pnwall   normal vector to wall faceN    number of nozzlesN P    number of ports for fluid devicem   Poisson’s ratio for diaphragm material p   pressure field of fluid

    P  p   average value of pressure   p over port  pP 0   ambient pressureP F   air pressure in air filling chamberP G   gas pressure upstream of solenoid valveQ   volume flow rate (used in two-port lumped

    fluid element)Q p   average flow rate over port prthresholdpressure ratio that marks that transition

    from sonic to subsonic flowRo   outer radius of throttleRi   inner radius of throttle

    RN   radius of nozzlesq   density of fluids   coordinate along mean stream line in flow

    in lumped fluid elementS (s) cross-sectional area of lumped fluid ele-

    ment at position  sS  p   area of port p  of fluid devicer   surface tension constant for fluidt   timeT   fluid stress fieldTvis   viscous contribution to fluid stress field

    T

    s   thickness of diaphragmH(Æ) unit step functionh   contact angle for fluid–air–wall interfacev   fluid velocity fieldvS    velocity of moving fluid surfacev p   velocity of fluid at port pV D(xD) volume displaced by diaphragm with cen-

    ter displacement  xDV Dispensed  volume of fluid dispensedxD   the displacement coordinate for the center

    of the diaphragm

    142   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    3/29

    In addition, if the modeling simulations and corresponding data correspond, then one can say that one under-stands the operation of the pump, to an appreciable degree. The operational parameters include drivepressure, throttle aperture, and the diaphragm displacement. Geometrical properties include the pressure headfrom the reservoir, the shape of the arrangement of the interior volumes and surfaces of the pump, as well asthe length and diameter of the nozzles. Other design considerations include, the number of dispensing nozzles,material characteristics, and very importantly, a nitrogen inerting system to limit the oxidation of the solder atall times during the operation of the pump. Given the complexity of the system, and the desire to predict tran-sient behavior, a comprehensive CFD model proved intractable. Thus, we were led to pursue a lumped-ele-ment type of analysis for the dispense operation. However, our lumped parameter analysis of the fluid flow

    excluded any detailed model of drop formation.

    v   gravitational potential functionz Z -coordinate of gravity field

    Z  p   average value of  z-coordinate over port  p

    Compressed

    Gas

    Solenoid

    valve

    Nozzles

    ReservoirSolder

    Throttle

    Diaphragm

    Stop

    Diaphragm

    FillingChamber

    Nitrogen

    Inerting

    Manifold

    Fig. 1a. Schematic of an early prototype molten solder dispenser.

    Thermal Couple

    Solder Reservoir

    Main Chamber

    Optical Probe for

    Diaphragm

    Displacement

    Compressed Gas

    Line

    Nozzles

     (underneath

    and at center)

    Pins for

    Attachment to

    Robot

    Inlet for Inerting

    Nitrogen

    Throttle

    Adjustment

    Heating Coil

    Diaphragm Stop

    Adjustment

    Nitrogen

    Inerting Diffuser

    Fig. 1b. Final prototype molten solder dispenser.

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   143

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    4/29

    1.1. Previous work 

    The basic goal of the solder dispensing technology was to directly place solder onto substrates and deviceleads without the need for printing. The technology was directly aimed at fine-pitch device applications, andsolder bumping for flip-chip applications. The trend towards smaller dimensions was one of the main drivers

    for this technology. The leads (say on a microprocessor device) may have a 400-lm pitch. Conventional solderprinting techniques are problematic at these dimensions, in part, because bridging may occur between theleads. Additionally, the promise of fluxless soldering, and reduced thermal stress, were also major consider-ations for pursuing the technology.

    There were other efforts and techniques, similar to the effort discussed in this paper, which developed work-ing prototypes. For example, the MPM Corporation developed two ‘‘metal jetting’’ technologies  [1,2]. Onetechnique involved an electrostatically guided stream of solder balls; a continuously dispensed stream of mol-ten solder was broken into a stream of balls through surface-tension induced instabilities. The stream of balls(50–300  lm in diameter) was directed onto electronic substrates with sub-micron placement accuracy. MPMalso developed a piezoelectrically actuated, single nozzle, ‘‘drop on demand’’ system, more akin to the systemdiscussed in this paper. This technique is essentially a copy of the inkjet dispensing scheme; it employs a cap-illary tube, one end of which is connected to a solder reservoir and the other end is a nozzle. The nozzle end

    has a surrounding piezoelectric element, which, when electrically pulsed, rapidly ejects the solder. The dynam-ics of the pump described in this paper is analogous to that used in the ink jet process, although at a firstglance this may not be apparent. In both the inkjet scheme, and in our scheme, the dispensing technique relieson a balancing of fluid momentum at the two ends of a capillary tube. This balancing requires a connection of one end of the tube to a reservoir with a free surface. The other end of the tube is dominated by capillaryeffects (surface tension forces).

    Another interesting technique was developed in Germany by Pac Tech-Packaging Technologies [3]. In thistechnology, preformed solid solder balls (80–760  lm diameters) are ejected from a capillary tube. As the ballsare ejected they are melted by a laser. What is interesting about this process is that the volume of the balls ispredetermined, and the laser not only melts the ball but can add extra heat to the balls in a controlled fashionso that they can better form intermetallic joints with the substrate.

    We mention in passing that, in addition to the mechanics of the dispensing process, the solidificationprocess and formation of the intermetallic joints, which the dispensed solder makes with the substrate and/or device leads, must also be controlled. A discussion of the modeling of the joint formation can be foundin [4–6]. It appears to the authors of this paper that there is more in the literature regarding joint formationthan dispense modeling, probably because of the proprietary nature of the workings of the dispensersthemselves.

    2. Description of the pump and its operation

    The pump is constructed almost entirely of titanium, so that the pump walls are not chemically reactivewith the molten solder. The pumping action is executed with a diaphragm driven by a controlled (open loop)actuation of pneumatic (solenoid) valves. In some applications, the pump is required to simultaneously dis-pense 256 drops of solder (where each drop is typically about 200  lm in diameter) onto the 4  ·  64 = 256 leadsof a microprocessor device. Nozzle arrays (64 nozzles in one array) were fabricated with silicon micro-machin-ing techniques. For dispensing a single drop, a single nozzle was employed. In this case, the single nozzles werefabricated in thin titanium plates with very fine drills, as small as several thousands of an inch in diameter. Thenozzles were typically 100 lm in diameter and 1000 lm long. The flow is, of course, highly viscous within thesechannels.

    In preparation for dispensing, one must prime the pump; i.e., one must fill the pump entirely with liquidsolder, with virtually no air trapped inside. Because the dispense volumes are so small, small amounts of trapped air can lead to an uncontrollable compressibility. Once the pump has been primed, it is in the restingstate, and one may wonder why the pump does not then leak. There are no valves preventing the solder fromdraining out the nozzles at the bottom of the pump. Indeed, the pump does not leak because the surface ten-

    sion of the solder–air interface in the nozzles supports the pressure head.

    144   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    5/29

    One may divide that part of the pump, which contains the fluid, into four sections: the solder reservoir,main pump chamber, the adjustable throttle between the reservoir and the main pump chamber, and the dis-pensing nozzle or nozzles. These four sections of fluid interact during the dispense process, which proceeds asfollows: the diaphragm pushes on the fluid in main pump chamber; the fluid then has two paths to flow. Onepath goes from the diaphragm up through the adjustable throttle valve, and into the pump reservoir. The

    other path from the diaphragm goes down to the nozzles. We have several control knobs to help tune the dis-pense process. We can control displacement and speed of the diaphragm, the pressure head in the reservoir,and the constriction of the throttle valve. The diaphragm displacement is limited by a ‘‘stop’’, which allows thediaphragm to move, typically, between one to two thousandths of an inch. Once the diaphragm hits the stop,the fluid may still flow through the throttle to the reservoir, or down through the nozzles, but only in onedirection. A proper selection of the throttle position will cause the fluid to flow up, creating a draw back mech-anism on the nozzle flow, which is adjustable and independent of the diaphragm drawback, which occurs wellafter each dispense. In the absence of the throttle-mitigated draw back, the pump may continue to dispense atlow velocities toward the end of the dispense cycle, resulting in misdirection and dribbling of the solder ema-nating from the nozzles. Conversely, excessive drawback can lead to high retraction velocities within the noz-zles, causing air to be drawn into the pump. The optimum control of drawback is therefore essential tosuccessful dispensing.

    To be able to perform a succeeding pumping action, one must allow the diaphragm to relax to its restingposition. We do not include this part of the process in our analytical description, since no fluid is dispensed.

    3. Analytical formulation

    From this characterization of the dispense process, one can see that a description of the dispensing processrequires the analysis of coupled processes from at least three physical domains. There is

    1. A pneumatic (air flow) process: the filling of the air chamber upstream of the diaphragm by opening andclosing a solenoid valve.

    2. A mechanical process: the dynamic response of the diaphragm due the applied pressure from the air pulse

    and back pressure of the fluid.3. A fluid (liquid) flow process: the flow of solder within the pump volume and out from the nozzles.

    To this end, we provide a representation for the air pulse that drives the diaphragm, we use a simple Hook’sLaw model for lumped-element representation of the diaphragm, and we obtain an analytical representation, alumped-element formulation, of the physics in the four fluid regions of the pump.

    We treat the driving function as an applied boundary condition, and the regions are connected togetherthrough the continuity of physical variables. A quantitative description of the pump behavior is then obtainedby simultaneous solution of the aforementioned analytical forms for a given set of design parameters.

    Before we construct the various lumped elements for the volume regions of the pump, we give below a gen-eral derivation of the approximate dynamics of the fluid lumped elements.

    3.1. Construction of lumped fluid elements – general treatment

    In the next two sections we develop a formalism to allow us to construct lumped-element models for thefour regions volumes of the pump in which the molten metal flows. In Section 3.2 below, we develop a generalBernoulli lumped fluid element. This element will be used for the main chamber and reservoir region volumes.For these regions, the salient engineering approximation that we make is that the fluid dynamics of the moltensolder is well approximated by potential flow. We argue that the fluid motion in the main chamber and res-ervoir is essentially inviscid because the significant dimensions are relatively large (the total volume in eitherregion is very much larger than the volume displaced in a single pulse), and the velocities are small (no morethan a few millimeters per second.) The sudden change in magnitude and, in some cases, direction of the flowafter the diaphragm hits the stop could create local vortices that might have a minor impact on the subsequent

    flow, but we neglect their impact in our analysis.

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   145

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    6/29

    In Section  3.3, we develop a fluid element applicable to flow in narrow tubes, where viscous effects areimportant. The treatment for these fluid elements is, in one sense, more general than that for the Bernoullifluid elements, in that we account for viscous effects. It is less general, in another sense, in that we must relyon specific known solutions to be able to complete the dynamics of the element. The reason that we need toinvoke these specific solutions is that there is no general solution for the velocity field within a control volume

    for the Navier–Stokes equation, as far as the authors know, in terms of just data on the surface of the controlvolume.

    3.2. General formulation for a Bernoulli lumped fluid element

    We begin our construction of Bernoulli lumped fluid elements by following Stevenson  [7]. We specify aninviscid flow of velocity  v  in a region  V   with boundary   oV . The fluid has density  q, pressure  p, and gravita-tional body force density   f  = g. The boundary of the region volume, or the control volume, is comprised of a ‘‘wall’’, through which no fluid passes, and a series of  N P  ports, through which fluid may pass, as illustratedby Fig. 2. One may write for the boundary

    oV    ¼ oV   ports oV   walls  ¼  N  P 

     p ¼1

     F  p   F wall:   ð1ÞFor inviscid flow, the velocity  v is derivable from a potential  u.

    v ¼ ru:   ð2Þ

    With conservation of mass for incompressible flow, i.e.,

    r v ¼ 0;   ð3Þ

    the potential satisfies Laplace’s equation

    r2u ¼ 0:   ð4Þ

    We may speak of the physics within the lump volume in terms of the language of bond graphs  [8]. There is athrough variable, for us this variable is volume flow rate, and a cross-variable, for us the velocity potential.From Stevenson et al., we know that we may write an approximate integral form of Laplace’s equation as

    U p  ¼X N  P 

    q

     M  pqQq; p  ¼ 1; . . . ; N  P :   ð5Þ

     f = g

    1 1 1 1 1Face F : S ,Q , ,K  Φ 

    Φ 2 2 2 2 2

    Face F : S ,Q , ,K  

     N N N N N Face F : S ,Q , ,K  

     N Wall F F F F V  ++++=∂   ...21

    Port 1

    Port 2

    Port P N 

    n N 

    n1

    n2

    Region Volume V 

    Φ 

    Fig. 2. A fluid device with N  p ports. Each port has a face  F , an area S , and an outward pointing normal  n. There is an average flow rate  Qout of the port in the direction  n. There is an average potential U  (for Bernoulli flow), as well as a  K  constant that characterizes the kinetic

    energy out of the port in terms of  Q.

    146   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    7/29

    Here, N P  is the number of ports associated with the control volume,  U p is the average of the potential u  over athe face  F  p  of port  p, and  Qq  is the average volume flow rate out of the port  p,

    Qq  

    Z  F  p 

    dS 

    S  p ðS  p n p  vÞ:   ð6Þ

    The symmetric matrix,  M = {M  pq}, is a set of connection coefficients, essentially geometrical quantities, thatrepresents a generalized impedance relation between the through and cross variables. For example, if the po-tential  u  represented a static electric potential, then the connection coefficients would represent an (inverse)capacitance matrix elements between potential and charge.

    We note that mass conservation equation yields, with the averaging process over ports, the relationX N q

    Qq  ¼ 0:   ð7Þ

    We now add dynamics to our lump formalism in the form of Bernoulli’s equation

    qou

    ot   þ

    1

    2qv2 þ p þ q gz ¼ 0:   ð8Þ

    To convert to a lumped approximation, we average Bernoulli’s equation over the face  F  pZ  F  p 

    dS 

    S  p q

    ou

    ot   þ

    1

    2qv2 þ p þ q gz ¼ 0

    ! q

    oU p 

    ot   þ

    1

    2qhv2i p  þ P  p  þ q gZ  p  ¼ 0:   ð9Þ

    The quantities  Z  p and  P  p are, respectively, the average  z-value, and the average pressure over the port  p.Consider the quantity hv2i p. We typically do not have detailed information on the flow fields at the ports,

    but we may postulate, or impose, a boundary condition here. We will specify that the velocity at the port isapproximately proportional to the average normal flow  Q p out of the port

    vj p  ! ðQ p =S  p Þn p ;   so that  1

    2qhv2i p  !

    1

    2q

     K  p 

    S 2 p Q2 p :   ð10Þ

    The dimensionless constant K  p is an adjustable, or empirically determined, constant that expresses our imper-fect knowledge of the actual flow at the port  p. We expect K  p to be less than or on the order of 1 (the average of the square is less than the square of the average).

    If we now substitute the impedance relation Eq.   (5), into the averaged Bernoulli equation, Eq.   (9), weobtain the lumped-element relation that we have been seeking

    qX N 

    q

    d

    dt ð M  pqQqÞ þ

    1

    2q

     K  p 

    S 2 p Q2 p  þ q gZ  p  þ P  p  ¼ 0;   p  ¼ 1; . . . ; N :   ð11Þ

    Here we have a set of coupled non-linear first order differential equations for the flow through each port. Wemust also include the mass conservation relation, Eq.  (7), with this set.

    3.3. General formulation for a viscous lumped element

    The flow in the throttle and nozzles is different from that in the pump body and reservoir, since the fluidvelocities are significantly higher, and correspondingly the characteristic dimensions are smaller than that of the main chamber of the pumb body. Viscous effects could be important during certain phases of the pulse. Inaddition, in certain periods during the dispense process, the solder meniscus in the nozzles moves. Thus, ourgeneral treatment must consider moving boundaries as well as viscous effects.

    Following the procedure outlined by Slattery   [9], we begin by briefly deriving the generalized integralmomentum balance equations. Initially, the region volume, or control volume, has the same general geometryas employed for the fluid element illustrated in Fig. 2; we then specialize to control volumes with more tubular

    shapes, more akin to the nozzle and throttle element, such as that illustrated in  Fig. 3.

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   147

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    8/29

    The momentum balance equation is

    o

    ot ðqvÞ þ r ðqvvÞ ¼ ð r TÞT þ qf ;   ð12Þ

    where we will assume that the body force density  f  is derivable from a potential, and that the density is con-stant, i.e.,

    f  ¼ rv;   where  v ¼ g z:   ð13Þ

    We shall make the conventional assumption that the stress tensor  T partitions into an isotropic pressure  p  andviscous stress  Tvis

    T ¼  p I þ Tvis:   ð14Þ

    We then integrate over our time-changing control volume  V , and apply the divergence theorem.

    Z V   dV    oot ðqvÞ ¼ Z oV   dS n ðqvvÞ þ Z oV   dS ðT nÞ þ Z oV   dS nðqvÞ:   ð15ÞHere,  n   is an outwardly pointing normal to the surface   oV  of  V . Next, we transform the remaining volumeintegral above by applying the transport theorem to the momentum density,  qv, i.e.Z 

    V  

    dV    o

    ot ðqvÞ ¼

      d

    dt 

    Z V  

    dV   qv

    Z oV  

    dS ðn vSÞqv:   ð16Þ

    The velocity   vS   is the velocity of the moving boundary. We combine the last two equations to yield thefollowing:

    d

    dt 

    Z V  

    dV   qv ¼

    Z oV  

    dS n ðv vSÞqv þ

    Z oV  

    dS ðT nÞ þ

    Z oV  

    dS nðqvÞ:   ð17Þ

    The reader is referred to Fig. 2 again to see that the boundary  oV  is partitioned into two types: ports and walls.In principle, the surface of the port could be moving; e.g., the bottom port of a nozzle is a moving free surface.We assume that walls do not move and, of course, no fluid crosses a wall. Thus, if there are a total of  N P  ports,

    d

    dt 

    Z V  

    dV   qv ¼ X N  P  p ¼1

    Z  F  p 

    dS n p  ðv vSÞqv þX N  P  p ¼1

    Z  F  p 

    dS ðT n p  þ n p qvÞ

    Z  F wall

    dS nwall ðv vSÞqv

    þ

    Z  F wall

    dS ðT nwall þ nwallqvÞ:   ð18Þ

    We now simplify the RHS of this equation. Consider the second term on the RHS. We assume that the viscousforces across the surface  F  p  of a port are negligible

    Tvis nj F  p  ! 0;   ð19Þ

    2 2 2 2 2: , , ,F P Q S Z  

    n1

    1 1 1 1 1: , , ,F P Q S Z  

    n2

    V, S wall

    vS 

    v

    12ˆl l 

    s

    S(s)

    ⊥v

    | |

    Fig. 3. Model for control volume V  used to derive viscous lumped elements.

    148   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    9/29

    given that the port surfaces are generally perpendicular to the flow. HenceZ  F  p 

    dS ðT n p  þ n p qvÞ ¼

    Z  F  p 

    dS Tvis n p  þ

    Z  F  p 

    dS n p ð p þ qvÞ ¼

    Z  F  p 

    dS n p ð p þ qvÞ:   ð20Þ

    For the third term on the RHS of Eq.  (18), the normal component of the fluid velocity  nwall Æ  v  vanishes at a

    wall as does  nwall Æ

     vS . HenceZ  F wall

    dS nwall ðv vSÞqv ! 0:   ð21Þ

    For the fourth term on the RHS of Eq. (18), we substitute our expression for  T, Eq. (14), and for v, Eq. (13), toobtain

    Gwall  

    Z  F wall

    dS ðTvis nwall nwallð p þ q gz ÞÞ:   ð22Þ

    Here, G wall is the force that the fluid exerts on the bounding wall. Our simplified integral form then reduces to

    d

    dt  Z V   dV   qv ¼ X N  P 

     p ¼1 Z  F  p  dS n p  ðv vSÞqv X N  P 

     p ¼1 Z  F  p  dS n p ð p þ q gz Þ þ Gwall:   ð23ÞThe equation above forms the basis for our integral momentum balances in the throttle and nozzles. Below wewill apply this equation to a more specific lump geometry than that of  Fig. 2. As with the Bernoulli lumpedelements, one must solve Eq. (23) in conjunction with the mass balance, Eq. (7).

    3.4. Application of the viscous integral momentum form to tubular elements

    We now tailor the integral momentum equation, Eq. (23), more to our application, with the simple controlvolume of  Fig. 3 in mind. This control volume has the basic shape of the throttle and nozzle volume regions.

    First we consider the volume term of the LHS of Eq. (23), and examine Fig. 3; we define l12 l̂  to point fromport 1 to port 2, and positive flow is from port 1 to port 2. We wish to relate this volume integral to the average

    volume flow rate Q. In general, one may split the velocity field, v, into parallel components, vk (along the axialdirection l̂ ), and perpendicular components,  v?. We then assume that the perpendicular components approx-imately average to  0, i.e.,

    d

    dt 

    Z V  

    dV   qv ¼  d

    dt 

    Z V  

    dV   qvk þ d

    dt 

    Z V  

    dV   qv?

     zfflfflfflfflfflffl}|fflfflfflfflfflffl{ 0:   ð24Þ

    We then express the remaining integral above in terms of the average volume flow rate, as follows:

    Z V   dV   vk  ¼ Z l d sS ð sÞ Z SdS 

    S ð sÞ

    vk  zfflfflfflfflfflfflfflfflffl ffl}|fflfflfflfflfflfflfflfflffl ffl{ vkð sÞ̂l 

    ¼ l12 Z ld s

    l

      S ð sÞvkð sÞ  zfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ Q

    l̂  ¼ l12Ql̂ :   ð25Þ

    From mass conservation we know that  Q  = Q2 = Q1, where Q2 is positive, going out of the second port. Thevolume term of Eq. (24) then reduces to

    d

    dt 

    Z V  

    dV   qv ! l̂ qdðl12QÞ

    dt   :   ð26Þ

    Next we consider the first surface term on the RHS of Eq.  (23). We may write this term as

    X2

     p ¼1

    Z  F  p 

    dS n p  ðv vSÞqv ¼

    qS 1

    Z  F 1

    dS 

    S 1ðn1 v1v1Þ þ qS 2

    Z  F 2

    dS 

    S 2ðn2 v2v2Þ;   F 1   and   F 2   fixed;

    qS 1 Z  F 1dS 

    S 1ðn1 v1v1Þ;   F 1   fixed;   F 2  moving with  v2  ¼ vS:

    8>>><>>>:

    ð27Þ

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   149

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    10/29

    This term is a boundary kinetic energy term, similar to that encountered in the Bernoulli lumped-element for-mulation. As with the Bernoulli term, we generally do not know the details of the flows at the entrances andexits of our simple control volumes. We observe that we have an average over a quadratic dependenceon velocity, at each port. Therefore, much as we did for the Bernoulli lumped element, we postulate akinetic boundary condition (based on a possibly empirically determined constant.) We say that, if   n p  l̂  ¼

    sgnðQ p Þ,Z  F  p 

    dS 

    S  p n p  ðv p v p Þ ! sgnðQ p Þ̂l 

      1

    2 K  p Q

    2 p =S 

    2 p 

    ;   p  ¼ 1; 2:   ð28Þ

    We write, with sgn(Q) = 1,X2 p ¼1

    Z  F  p 

    dS n p  ðv vSÞqv ¼ l̂   1

    2q

     K 2ðvSÞ

    S 2Q2

    1

    2q

     K 1

    S 1Q2

    :   ð29Þ

    To complete our construction of a viscous fluid element, we move to the second surface term in Eq.  (23) andexpress this surface integral as

    X2 p ¼1 Z  F  p  dS n

     p ð p þ q gz Þ ¼ S 1 Z  F 1dS 

    S 1 n1ð p þ q gz Þ þ S 2 Z  F 2dS 

    S 2 n2ð p þ q gz Þ

    ! l̂ ðS 2 P 2 S 1 P 1Þ þ l̂ q g ðS 2Z 2 S 1Z 1Þ:   ð30Þ

    Finally, we consider the wall term Gwall, which is the force that the fluid exerts on the wall through viscous andpressure forces. We consider only those components along our direction of flow   l̂ , in which case, we pick off primarily viscous forces, since   l̂ 12  tends to be orthogonal to the normal vector, the direction of the pressureforces. So we write

    l̂  Gwall  G ðQÞ;   ð31aÞ

    where  G (Q) is  the force of the wall on the fluid . We allow for the dependence of  G  on fluid velocity through adependence of  G  on  Q. Note, also, that a provision must be made for  G  to switch signs, if the flow direction

    changes direction. Thus, G (Q) must have the property thatG ðQÞ ¼ sgnðQÞjG ðQÞj:   ð31bÞ

    We now collect all terms together, and ‘‘dot’’ all terms in the vector equation, Eq.  (23) by   l̂ , to remove thevector dependence. The dynamic equation for our tubular viscous lumped element is then

    q d

    dt ðl12QÞ ¼

    1

    2q

      K 1

    S 1

     K 2

    S 2

    Q2 þ q g ðS 1Z 1 S 2Z 2Þ G ðQÞ þ ðS 1 P 1 S 2 P 2Þ:   ð32Þ

    As with the Bernoulli equation, the dynamical equation above is not dependent on our chosen reference port.If one changes the sign of  Q, and also interchanges the port indices, the equation remains invariant.

    3.5. Lossy port condition for viscous lumped element

    In our application, we have viscous lumped elements, such as the throttle, which adjoin ‘‘Bernoulli’’ lumpedelements such as the main chamber and the reservoir. However, in our formulation so far, we have notaccounted for the pressure losses that can occur at a port for flow from a constricted region to one more open.We will do this accounting by constructing a simple viscous lumped-element model of the transition region atthe port and then affix this element to the viscous lumped element (see  Fig. 4).

    We will assume in the transition region, that if flow is from the viscous region to the Bernoulli region, thenwe have an adjustable loss (likely, almost total loss) of the kinetic energy. This loss leads to a pressure change.If the flow is in the reverse direction, we then assume little loss, again adjustable. We then combine the tran-sition region with the viscous region, through the continuity of pressure, to form a single region. We shrink

    this combined region back to the size of the original viscous region.

    150   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    11/29

    Referring to Fig. 4, and using Eq.  (32), we write the lumped equation for the transition region

    q  ddt 

      l0Qð Þ ¼ 12q   K 

    002

    S 002 K 

    002

    S 002

    Q2 þ q g S 02Z 

    02 S 

    002 Z 

    002

    G 0ðQÞ þ ðS 02 P 

    02 S 

    002 P 

    002Þ:   ð33Þ

    The model that we use for the port loss, as mentioned above, is to assume a pressure drop,  DP viscous, that arisesfrom the viscous forces originating locally at the port wall. We then say,

    D P viscous ¼G 0ðQÞ

    S 2!

    1

    2q K loss2   ðQÞ

    Q2

    S 22:   ð34Þ

    We define,

     K loss2   ðQÞ þk 

    out;   Q >  0;k in;   Q <  0;

      !ideally   1;   Q >  0;0;   Q <  0;

      ¼ HðQÞ:   ð35ÞNext, we write the corresponding equation for the lower viscous region

    q d

    dt ðlQÞ ¼

    1

    2q

      K 1

    S 1

     K 02S 02

    Q2 þ q g ðS 1Z 1 S 

    02Z 

    02Þ G ðQÞ þ ðS 1 P 1 S 

    02 P 

    02Þ ð36Þ

    We wish to eliminate the intermediate pressure  P 02  from the equations for the transition region, Eq.  (33), andthe lower viscous region, Eq.  (36). This elimination is easily accomplished, if we just add the two equationstogether; we obtain

    q d

    dt   ðl þ l0ÞQ½ ¼

    1

    2q

      K 1

    S 1

     K 002S 002

     K loss2   ðQÞ

    S 2

    Q2 þ q g ðS 1Z 1 S 

    002Z 

    002Þ G ðQÞ þ ðS 1 P 1 S 

    002 P 

    002Þ:   ð37Þ

    We then shrink the transition layer, so that

    l0 ! 0;   so that  S 002  ! S 2;   Z 002  ! Z 2;   K 

    002  !  K 2;   and  P 

    002  !  P 2:   ð38Þ

    We find the final form for the viscous lumped element, with a modified kinetic boundary condition to accountfor port loss, is

    q d

    dt ðl12QÞ ¼

    1

    2q

      K 1

    S 1

     K 2 þ K loss2   ðQÞ

    S 2 Q2 þ q g ðS 1Z 1 S 2Z 2Þ G ðQÞ þ ðS 1 P 1 S 2 P 2Þ:   ð39Þ

    Fig. 4. Model used to develop boundary condition for a port that transitions between a viscous region and a Bernoulli region.

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   151

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    12/29

    3.6. Comparison of viscous and Bernoulli fluid elements

    The similarity in form between the viscous fluid element equation, and that of the Bernoulli, which wereproduce below

    qX N 

    q

    d

    dt ð M  pqQqÞ ¼

    1

    2q

     K  p 

    S 2 p  Q

    2

     p  q gZ  p   P  p ;   p  ¼ 1;. . .

    ; N ;   ð40Þ

    is apparent from inspection. However, we can make the similarity even more apparent, if we consider applyingthe Bernoulli lumped-element equation to a two-port device, with a port 1 and a port 2. We consider port 1 thereference port; in this case, we subtract the Bernoulli lumped equation for port 1

    q d

    dt ð M 11Q1 þ M 12Q2Þ ¼

    1

    2q

     K  p 

    S 2 p Q21 q gZ 1  P 1;   ð41Þ

    from that of port 2,

    q d

    dt ð M 21Q1 þ M 22Q2Þ ¼

    1

    2q

     K  p 

    S 2 p Q22 q gZ 2  P 2;   ð42Þ

    we use the fact that  Q1 + Q2 = 0, define,  Q2 Q, in addition, define

     M eff 12    M 11 2 M 12 þ M 22  ¼ trðMÞ 2 M 12  ¼  M eff 21 :   ð43Þ

    We arrive at the Bernoulli lumped-element equation for a two-port element,

    q d

    dt ð M eff 12 QÞ ¼

    1

    2q

      K 1

    S 21

     K 2

    S 22

    !Q2 þ q g ðZ 1 Z 2Þ þ ð P 1  P 2Þ:   ð44Þ

    Note that the equation is invariant, if we let  Q ! Q, and switch port labels. We see that the Bernoulli for-mulation and the viscous formulation give rise to similar lumped-element representations, which one mightexpect, but, of course, the Bernoulli element does not contain the viscous force contribution. The similarityis closer if one identifies the connection coefficients, the M ’s, as essentially inverse characteristic length scales,

    approximately the length along the center fluid stream line, divided by an area M eff 12   $ l=S :   ð45Þ

    Of course, this relationship is essentially that found for the inverse of the capacitance between two chargedplates in electrostatics.

    4. Derivation of lumped elements for the pump volume regions

    In the next six subsections, we present a representation for the driving pneumatic pulse, and give specificrepresentations of the five lumped-element volume regions for the pump.

    4.1. Pneumatic driving force

    In this section, we provide an analytical function that describes the time dependence of the build-up of thepressure, P F(t), within the air filling chamber, by flow through a solenoid valve. This pressure drives the dis-placement of the diaphragm (see Fig. 5a); this pressure function provides a boundary condition for the lumpedelement of the diaphragm.

    The formation of the transient pressure pulse is well described by the filling of a hot chamber with a coldgas through an orifice. The filling of the chamber proceeds as follows: If, at the beginning of the pulse, thepressure ratio (the pressure downstream of the orifice divided by the pressure upstream of the orifice,   P F/P G, is less than   rthreshold = 0.526  [10]), we initially have choked (sonic) flow through the solenoid valve. Asthe chamber fills the downstream pressure rises, and, if the pulse is long enough, as is generally the case inour application, the pressure ratio,   P F/P G, rises above   rthreshold. The flow then becomes subsonic through

    the orifice. If the filling process were to continue indefinitely, the downstream pressure would rise enough

    152   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    13/29

    for the flow in the orifice to become incompressible. In our application, the pulse is not long enough for theincompressible state to be reached. If the initial pressure ratio is greater than rthreshold, then the ‘‘choked con-dition’’ never prevails; only the subsonic analysis is considered.

    Since the volume displacement of the diaphragm is small compared with the volume of the pneumaticchamber, the chamber volume may be assumed constant during the filling operation. Simultaneous massand energy balances for the chamber yield a relationship for the pressure transient in terms of the mass flow

    rate through the orifice, and the incoming gas temperature. The controlling resistance to the air flow lies in thesolenoid. Andersen [10] reports formulas for the mass flow rates through the orifice in terms of the drivingpressures, and gas temperature, for both subsonic and sonic flow in the orifice. These formulas are substitutedinto the aforementioned pressure transient relationship. The resulting equations that define the pressure tran-sient in the chamber for subsonic and sonic flow in the orifice are summarized below for convenience.

    For sonic flow into the hot filling chamber at temperature  T F and an initial pressure  P G (on the upstreamside of the solenoid valve), we have the pressure  P F   in the chamber versus time given by

     P Fðt Þ

     P G¼

      P 0

     P Gþ ðasonict Þ;   ð46aÞ

    where

    asonic 0:231c p S SC S

    V   F

     ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiT Fmmol R

    r   and   P 

    F

     P G6 r threshold:   ð46bÞ

    The filling gas is characterized by a heat capacity at constant pressure,  c p, and a molecular weight  mmol; thesolenoid valve has the phenomenological parameters S S, the effective cross-sectional area, and  C S, the orificedischarge coefficient.

    When the pressure in the filling chamber   P F/P G  increases above   rthreshold, the flow through the solenoidvalve becomes subsonic, then the pressure in the filling chamber versus time becomes

     P Fðt Þ

     P G¼ sin arcsin

      P 0

     P G

    þ ðasubsonict Þ

    ;   ð47aÞ

    where

    asubsonic ¼0:24c p S SC S

    V   F

     ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiT Fmmol

     R

    r   ;   and

      P F

     P G> r threshold:   ð47bÞ

    Once the flow becomes incompressible, Eq. (47a) is no longer applicable. In particular, Eq. (47a) can no longerbe used when the argument of the sine function exceeds  p/2. However, the relationship is accurate for therange of pulse times used in this application.

    4.2. Lumped element for the diaphragm

    The diaphragm is acted upon by the pressure difference across its two faces, as illustrated in  Fig. 5b. Thereis the driving air pressure from the filling chamber   P F  on one side, and the back pressure of the fluid   P D

    against the diaphragm on the other. We choose to model the equation of motion for the diaphragm in terms

    Solenoid Valve

    PG C S 

    PF 

    V F 

    S S 

    Air Filling

    Chamber

    Diaphragm

    Fig. 5a. Parameters used to model the compressed air pulse pressure.

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   153

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    14/29

    of the static displacement xD of the center of the diaphragm, essentially Hook’s Law. We will ignore the inertiaof the diaphragm in its dynamics. We know that Roark  [11] for a thin, clamped diaphragm of area,  S D, thecenter displacement, in term of the pressure differential across the diaphragm, is

    k D xD ¼ ð P F  P DÞS D;   ð48Þ

    where

    k D  ¼  E 

    ð1 m2Þ16s3

    d 2Dð49Þ

    The constants E  and  m, are the modulus of elasticity and Poisson’s ratio, respectively, of the diaphragm mate-rial (in our case titanium.) The constants  s  and  d D are the diaphragm thickness (on the order of .010 in.) anddiameter (on the order of 1.00 in.), respectively. Again, using the formulas provided by Roark  [11], we canrelate the center displacement,  xD, to the volume,  V D(xD), displaced by the diaphragm as

    V   D ð xDÞ ¼ 1=3S D xD:   ð50Þ

    The volume flow rate generated by the diaphragm is then

    QD  ¼  d

    dt 

    V   D ð xÞ ¼ 1=3S Dd xD

    dt 

      ;   or  d xD

    dt 

      ¼ 3QD

    S D

    :   ð51Þ

    The maximum displacement of the diaphragm is  DxD.

    4.3. Lumped element for the main chamber

    We next apply the Bernoulli lumped element, developed in Sections 3.2 and 3.5, to the flow of solder in themain body. There are two distinct phases of pump operation that need to be analyzed when describing thedynamics of the flow in the pump as a whole.

    A first phase, which we will call the driving phase, begins when the solenoid valve opens to allow com-pressed gas into the pneumatic chamber upstream of the diaphragm. As the chamber fills, the diaphragmdeforms, only slightly, due to pressure build-up on the pneumatic side; the diaphragm displaces molten solderin the body of the pump. During this phase, the flow in the pump body splits into two streams, one leads to thereservoir, and the other to the dispensing nozzles, as shown in Fig. 5c. The solder meniscus in the nozzles, seeFig. 5f , begins its displacement from its resting position at the top entrance to the nozzles. The fluid motion istherefore generally downward in the nozzles, and generally upward in the throttle and the reservoir. The firstphase persists until the diaphragm hits the stop, which controls its maximum displacement; its motion haltsvirtually instantaneously. If we ignore the possibility of cavitation, then this occurrence has the effect of instan-taneously coupling the two streams, which were previously independent.

    In a second phase, which we will call the retraction phase, and which begins after the diaphragm hits thestop, the capillary forces in the nozzles work against inertial, gravitational, and viscous forces to eventuallyforce the solder upward, until the solder interfaces rise to the top of the channels. A single effective streamis sufficient to describe the flow in the pump body during this period. The flow associated with this streamcould be either downward or upward immediately after the diaphragm impact, depending on the throttle

    and nozzle parameters.

    Main

    Chamber

    Diaphragm

    Stop

    Diaphragm Parameters

    k  D, S  D

     x  D

    S  D

    PF 

    P D

    ∆ x  D

    Fig. 5b. Parameters used to model diaphragm physics.

    154   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    15/29

    In light of the preceding discussion, we will develop lumped elements for the main chamber that describethe transient evolution of the system during the two distinct phases.

    4.4. Flow in the main chamber – driving phase

    We apply the two-port lumped Bernoulli equation, Eq. (44), to the flow in the main chamber. During thedriving phase of the dispense process we shall ignore the mutual interactions between the ports correspondingto the throttle entrance and the nozzle entrances. As discussed earlier, the treatment is equivalent to appor-tioning the pump body into two two-port devices, each beginning at the diaphragm port  D, which we useas the reference port. In our treatment, we divide the diaphragm into two effective ports, port  Dtop and portDbot. The line that divides the diaphragm diameter  d D between its ‘‘top’’ part and its ‘‘bot’’ is not necessarily inthe middle of the diaphragm, but we will represent it as so. The resulting balances, in the driving phase, are asfollows. For the path from the diaphragm to the bottom of the throttle,   l DT,

    Diaphragm (port  Dtop) to Throttle (port  T bot)

    q d

    dt ½ M eff DTQT ¼

    1

    2q

      K topD

    ðS topD   Þ2

     K botT

    S 2T

    !Q2T þ q g ðZ 

    topD   Z 

    botT   Þ þ ð P D  P 

    botT   Þ:   ð52Þ

    For the path from the diaphragm to the nozzles,  lDNtop , we treat all of the nozzles as identical and independent,so that they combine into one effective port with total flow QtotalN   ¼  NQN, and total port area S 

    totalN   ¼  NS N; thus

    Diaphragm (port  Dbot) to Nozzles (port  N top):

    q d

    dt ½ M eff DNð NQNÞ ¼

    1

    2q

      K botD

    ðS bot

    D   Þ

    2

      K topN

    ð NS NÞ

    2 !ð NQNÞ2 þ q g ðZ botD   Z 

    topN   Þ þ ð P D  P 

    topN   Þ:   ð53Þ

    We have chosen positive   QT  to be up, diaphragm to throttle, and positive   QN   to be down, diaphragm tonozzles.

    Even though the diaphragm is not necessarily divided evenly, and in fact the division dynamically shifts, theeffect on the overall dynamics is negligible. We argue that since K topD   , K 

    botT   , K 

    botD   , and K 

    topN   are all of order 1, and

    that, typically

    S T   S botD   ;   and  NS N   S 

    topD   :   ð54Þ

    then

     K topD

    ðS topD   Þ

     K botT

    S 2T

    ;   and  K botD

    ðS botD   Þ

      K topN

    ð NS NÞ2 :   ð55Þ

    , , ,top top top

     D D D DP S K Z  

    , , ,bot bot bot  

     D D D DP S K Z  

    ( ) ( ), , , ,top top top N N N N N P N S K Z N Q

    Port D, S  D Q D

    h DT 

    Port T bot 

    Port N top

    , , , ,bot bot bot  

    T T T T T  P S K Z Q

    l  DT 

     top DN l 

    l  NT 

    d  D

    top DN h

    h NT 

    Fig. 5c. Parameters used to characterize lumped element for the main pump chamber.

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   155

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    16/29

    Thus, we will ignore the terms  K topD   =ðS topD   Þ

    2, and  K botD   =ðS 

    botD   Þ

    2. In addition, if we let

    Z topD   Z 

    botT   ¼ ðhDT d D=4Þ;   ð56aÞ

    and

    Z botD   Z topN   ¼ ðhDNtop  d D=4Þ;   ð56bÞ

    then we may rewrite the lump equations for the driving phase in the main chamber as

    q d

    dt ½ M eff DTQT ¼

    1

    2q

      K botT

    S 2T

    !Q2T q g ðhDT d D=4Þ þ ð P D  P 

    botT   Þ;   ð57Þ

    and

    q d

    dt ð NM eff DNQNÞ ¼

    1

    2q

      K topN

    S 2N

    !Q2N þ q g ðhDNtop  d D=4Þ þ ð P D  P 

    topN   Þ:   ð58Þ

    We perform one last manipulation on the equations above, to put them in a form that will be a bit more usefulfor us. We let

    S T M eff DT  ¼ lDT  ¼ const:   ð59Þ

    Similarly, we let

    ð NS NÞ M eff DT  ¼ lDNtop  ¼ const:   ð60Þ

    The parameters  l DT and  lDNtop  are characteristic length scales in the pump body associated with the streamsfrom the diaphragm to the throttle, and from the diaphragm to the nozzle, respectively.

    Our new equations are then

    qlDT

    S T

    d

    dt QT  ¼

    1

    2q

      K botT

    S 2T

    !Q2T q g ðhDT d D=4Þ þ ð P D  P 

    botT   Þ;   ð61Þ

    and

    qlDNtop

    S N

    d

    dt QN ¼

    1

    2q

      K topN

    S 2N

    !Q2N þ q g ðhDNtop  d D=4Þ þ ð P D  P 

    topN   Þ:   ð62Þ

    One can see from these equations, with the attendant approximations used to derive them, that we have re-moved reference to the detailed position of dividing line across the diaphragm for the two flows in the mainchamber of the pump, and thus effectively decoupled the two streams.

    Of course, the mass balance is still enforced. We must take the volume flow rate of the diaphragm QD to benegative, for flow into main chamber. Conservation of mass for the main chamber is then

    QD þ QT þ NQN ¼ 0:   ð63Þ

    4.5. Flow in the main chamber – retraction phase

    The flow in the pump body after the diaphragm is halted is significantly different from the flow during for-ward diaphragm motion. It becomes a single two-port device, with fluid moving from the nozzles to the throt-tle or vice versa. The change in flow pattern implies that we need a single characteristic length scalerepresenting the interaction between the two ports. We shall treat the nozzles as the reference port. Eq.(44) then implies that

    Nozzle (port  N top) to Throttle (port  T bot):

    q d

    dt 

    ð M eff NTQTÞ ¼1

    2

    q  K 

    topN

    ð NS NÞ2

     K botT

    S 2

    T !Q2T þ q g ðZ 

    topN   Z 

    botT   Þ þ ð P 

    topN    P 

    botT   Þ ð64Þ

    156   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    17/29

    We define

    lNT   S T M eff NT;   and  hNT   Z 

    botT   Z 

    topN   ;   ð65Þ

    we arrive at the desired result

    qlNT

    S T

    ddt 

    QT  ¼12q

      K topN

    ð NS NÞ2

     K botT

    S 2T

    !Q2T q ghNT þ ð P 

    topN    P 

    botT   Þ:   ð66Þ

    The mass balance equation for the retraction phase in the main chamber is

    þ QT þ NQN ¼ 0:   ð67Þ

    4.6. The lumped element for the throttle

    We treat the throttle as an annular conduit of a known inner radius, Ri , outer radius, Ro, and height hT (seeFig. 5d). The amount of throttle is adjusted by moving a thin, slightly tapered, rod into the throttle neck, effec-tively increasing or decreasing the radius of the inner cylinder.

    If we now employ our general lumped model for a viscous tubular element, but with two lossy ports, we canwrite the equation of motion for the throttle as

    Throttle (port T bot) to Throttle (port  T top):

    q lT

    S T

    d

    dt QT  ¼

    1

    2q

      K botT   þ K botlossT   ðQTÞ

    S 2T

    !Q2T

    1

    2q

      K topT   þ K 

    toplossT   ðQTÞ

    S 2T

    !Q2T q ghT

    G TðQTÞ

    S T

    þ ð P botT    P topT   Þ:   ð68Þ

    We have added a second loss term to Eq.  (39) for the bottom throttle port; but, because the positive flow  QT isinto the bottom port, instead of out of the port, we must account for a sign change in  QT in the argument of  K botlossT   .

    As a model for the frictional term  G T (QT), we employ the standard steady state treatment of flow in anannulus Slattery [9, p. 76]. In this case

    G TðQTÞ ¼8

    pl

      lT

     R2ok

    QT;   ð69aÞ

    where

    k ¼ 1 j4 þ 1 j2lnðjÞ

      ;   and  j ¼  Ri R0

    :   ð69bÞ

     Ri

    hT  =lT 

     Ro

    Porttop

    T top

    T P  , S T  ,

    top

    T K   ,

    top

    T  Z 

    Portbot 

    T bot 

    T P  , S T  ,

    top

    T K   ,

    bot 

    T  Z 

    Viscosity µ

    Fig. 5d. Parameters used to model the annular throttle element.

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   157

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    18/29

    4.7. The lumped element for the reservoir

    We apply the lumped Bernoulli equation, Eq. (44) to the reservoir. In this case, the top of the reservoir is afree surface.

    We have (see Fig. 5e)Throttle (port  T top) to Reservoir (port  R):

    q ddt 

    ð M eff TRQRÞ ¼12q

      K topT

    S 2T

     K R

    S 2R

    !Q2R þ q g ðZ 

    topT   Z RÞ þ ð P 

    topT    P RÞ:   ð70Þ

    Instead of using the volume flow rate QR of the second port, we shall this time use the flow rate  QT. However,we have chosen QT to be positive flowing out of the main chamber, so we must attach a negative sign to QT inthe mass balance for the reservoir

    QT þ QR  ¼ 0:   ð71Þ

    In a similar manner to our previous approximations for the main chamber lumped element, we argue that

    S 2R   S 2T;   or that K 

    topT   =S 

    2T    K R=S 

    2R;   ð72Þ

    for our reservoir lumped element. In addition, we specify that

    hR   Z R Z topT   ;   and  lTR  S T M 

    eff TR:   ð73Þ

    In addition, at the top surface of the reservoir, the total pressure  P R is the sum of that due to the surface ten-sion force, and the reference atmospheric pressure,  P 0, or possibly an applied ‘‘bias’’ pressure  P B P 0. Weneglect the pressure arising from the surface tension force since the area of the top surface is relatively large.Thus, we take,  P R = (P B P 0) +  P 0. We let   l R be the characteristic length scale for the reservoir. Thus, ourlumped equation for the reservoir is

    qdðlRQTÞ

    dt   ¼

    1

    2q

     K topT

    S TQ2T q ghR þ ð P 

    topT    P BÞ:   ð74Þ

    We make one additional approximation. We observe that the free surface of the reservoir will move impercep-

    tibly during any phase of the dispense; i.e., we will assume that d l R/dt ! 0. In this case, our final form for thereservoir Bernoulli lumped element is

    q lR

    S T

    dQTdt 

      ¼1

    2q

      K topT

    S 2T

    !Q2T q ghR þ ð P 

    topT    P BÞ:   ð75Þ

    This equation is valid for either phase of the dispense process.

    4.8. The lumped element for a nozzle

    Fig. 5f  gives the basic picture of the flow of solder in the nozzles. The surface tension in the solder causes ameniscus to form around a moving contact line. At this contact line, there is always an upward force on the

    column of solder above.

    h R = l R

    PT  , S  R , K  R , Z  R ,Port R

    Port T top top

    T P  , S T  ,top

    T K   ,top

    T  Z 

    l R

    Fig. 5e. Parameters used in the description of the lumped element for the reservoir.

    158   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    19/29

    The treatment of the flow of the solder in the nozzles, which have length,   l N, and diameter   DN = 2RN,requires some finesse. We wish to describe the dispensing process accurately enough to predict the amountof solder dispensed, but at the same time we do not wish to tackle the very difficult dynamical problem of the instabilities that must develop to break the solder from the nozzles. (There are many papers on this topic.Several references include Chaudhary [13–15], Jeggers  [16,17].)

    With this philosophy in mind, we can qualitatively describe the motion of the solder column through thechannels as follows: During the driving phase, the solder meniscus in any particular nozzle moves downwardfrom the top entrance of the nozzle, as the diaphragm begins to move forward. One port for the nozzle is thusa free surface and a moving boundary. The meniscus quickly reaches the nozzle exit, and solder dispensing

    begins. The exiting solder stream necks, and detaches roughly two nozzle diameters beyond the exit. The dia-phragm hits the stop after approximately a millisecond or two time span, and the retraction phase begins; butthe downward fluid flow in the nozzle does not necessarily cease at this time. The amount of solder dispensedis counted (with some arbitrariness) as the amount of fluid that passes through a nozzle exit up to that timethat the velocity of the solder column vanishes. After the column stops, the flow in the nozzles soon reverses,the meniscus moves upward from the nozzle exit to the top. This retraction occurs under the action of surfacetension forces in the absence of the driving diaphragm motion.

    One may estimate the maximum error associated with the definition of the term, ‘‘amount of solder dis-pensed’’. From high speed video footage, the quantity of residual solder that draws back into the nozzles isat most that of a conical sliver attached to the nozzle exit (see  Fig. 5f (iii)) of height  hsliver 2DN. For nozzles100 lm in diameter, an estimate of the maximum error is given by, Estimated Dispense Error 1

    3p

     D2N

    4  hsliver !

    0:13 mm3

    , i.e. roughly 8–20% of the volume dispensed for data reported in this work. This estimate is actuallyquite conservative. In practice, the error is generally less than the estimate, since the conical sliver at break off is normally based in the interior of the nozzle by about one nozzle diameter, rather than at its exit.

    We again apply our equation for the tubular viscous lumped element, with a lossy boundary condition onthe top port, and moving boundary on the bottom port. Our reference port is the top of the nozzle. Hence, ourlumped element for the nozzles is (refer to Fig. 5f )

    q d

    dt ðlNQNÞ ¼

    1

    2q

      K topN   þ K 

    lossN   ðQNÞ

    S N

    Q2N þ S Nq ghN G NðQNÞ þ G R þ S Nð P 

    topN    P 0Þ:   ð76Þ

    We take the kinetic energy boundary term to vanish for the bottom port, since it is a moving boundary. Thereis the viscous force  G N (QN), the sign of which depends of the direction of flow. This term will oppose the flow,

    regardless of the flow direction, since it depends directly on  QN. We have Slattery [9, p. 77]

    2

    residual

     N 

    h

     D≈

    i) Meniscus

    Resting Position

     N l

     R N 

    h N   = l N 

    ii) Meniscus

    Moving in channel in

    the driving and

    retraction phases

    Porttop

     N 

    Q N  , S  N ,top

     N K    ,top

     N P   ,top

     N  Z 

    θ

    σ

    Contact

    Line

    Portbot 

     N 

    S  N ,   0bot 

     N K  =   ,   0P   ,bot 

     N  Z 

    Viscosity µ

    Surface

    Tension σ

    Contact

    Angle θ

    µ

     N  D

    iii) Dispense phase

    Fig. 5f. Three configurations of the nozzle meniscus and the parameters used to model the annular nozzle element.

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   159

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    20/29

    G NðQNÞ ¼8

    pl

      lNS N

     R4N

    QN:   ð77Þ

    There is also the capillary force  G R, or the surface tension force Batchelor  [12],

    G R  ¼ 2p RNr cosðhÞ:   ð78Þ

    The parameter r  is the surface tension. The surface tension force is applied along the contact line of the solid– liquid–air interface. (See Fig. 5f ) Because the nozzle walls are non-wetting, the contact angle is larger than 90;the force is always directed upwardly

    One may take the term dl N/dt   as the velocity of the meniscus. This moving boundary term then takes aform that is very similar to the kinetic energy term. Since

    dlNdt 

      QN  !Q2NS N

    :   ð79Þ

    If we now collect all the terms together, we have for the nozzle lumped element

    q lN

    S N

    d

    dt QN  ¼ q

      1

    S 2

    N

    Q2N þ1

    2q

      K topN   þ K 

    lossN   ðQNÞ

    S 2

    N !Q2N þ q ghN þ

    G R G NðQNÞ

    S Nþ ð P topN    P 0Þ:   ð80Þ

    To connect the nozzle volume flow rate  QN to the dispensed volume, we employ the mass balance equation forthe nozzles. By definition, the volume rate  QN  in the nozzles is

    QN  ¼dV   N

    dt   ;   ð81Þ

    We then say that the volume dispensed,  V Dispensed, is that volume that exits all nozzles during the time whenthe variable   l N  is greater than the length of the nozzle length  LN  to the time when the volume flow rate  QNvanishes (changes from positive to negative)

    V   Dispensed  ¼ Z   time when   QN¼0

    time when   lN> LN

    dt ð NQNÞ:   ð82Þ

    4.9. Assembly of lumped elements for the driving phase

    We assemble our equations for the driving phase of the pumping action. We have two paths, one from thediaphragm to the reservoir, D ! R, where we will thus assemble Eqs. (61), (68), and (75), and a second path,D ! N, from the diaphragm to the end of the nozzles, where we assemble Eqs. (62) and (80). To assemble theequations means to eliminate the unknown intermediate pressures and kinetic boundary terms between thelumps. We affect this assembly by just adding the equations together so that the intermediate pressure andkinetic terms pair-wise cancel. However, the port loss terms survive. We then obtain two equations, one equa-tion for the flow  QT, and another for the flow  QN. The resultant equation for the flow  QT  is

    q lDRS T

    ddt 

    QT  ¼ 12q K 

    drvT   ðQTÞ

    S 2TQ2T q ghDR G 

    TðQTÞS T

    þ ð P D  P BÞ;   ð83aÞ

    where we define,

    lDR   lDT þ lT þ lR;   ð83bÞ

    hDR   hDT d D=4 þ hT þ hR;

     K drvT   ðQTÞ  K botlossT   ðQTÞ þ K 

    toplossT   ðQTÞ:   ð83cÞ

    If assemble Eqs. (62) and (80), we find the resultant equation for the flow  QN, for the flow path  D ! N ,

    qlDN

    S N

    d

    dt 

    QN ¼1

    2

    q  K drvN   ðQNÞ

    S 2

    N !Q2N þ q ghND þ

    G R G NðQNÞ

    S N

    þ ð P D  P 0Þ;   ð84aÞ

    160   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    21/29

    where we define,

    lDN   lDNtop þ lN;   ð84bÞ

    hND   hN þ ðhDNtop  d D=4Þ;   and   ð84cÞ

     K drvN   ðQNÞ  K lossN   ðQNÞ 2:   ð84dÞ

    These equations are driven by the diaphragm pressure,P D, which is subsequently driven by the air pulse gaugepressure  P F, i.e., from Eq. (48a)

     P D ¼  P F k D xD=S D:   ð85Þ

    To close the set of equations for the driving phase, one must include the relationship between  xD, and the flowrate  QD  from Eq. (51). Hence,

    d xDdt 

      ¼ 3QDS D

    ¼  3

    S DðQT þ NQNÞ:   ð86Þ

    For the driving phase, we have a system of three coupled first order non-linear differential equations, Eqs.(83a), (84a), and (86)  for the three unknowns,  QT,  QN, and  xD.

    We remark that the equations for the individual flows are quite transparent in their meaning. They lookvery much like that for our general tubular viscous element. Under the assumption of independent flows,one up through the throttle, and the other down through the nozzles, one may an essence write down theseequations for the flows   QT, and   QN, by inspection, using an effective cumulative path length, say   l DR, forexample. One includes the pressure boundary conditions at the beginning and end of the path, as well asany viscous losses along the path, and finally a pressure head that results from the vertical change in heightfrom the beginning port to the ending port.

    4.10. Assembly of lumped elements for the retraction phase

    To assemble the equations for the retraction phase, we add the dynamical equations along the path N ! R,using Eqs. (80), (66), (68), and (75). In addition, we must employ the continuity condition for the retractionphase, namely

    QT  ¼  NQN;   QD ¼ 0:   ð87aÞ

    The assembled dynamic equation is then,

    qlNR

    S N

    d

    dt QN ¼

    1

    2q K retðQNÞQ

    2N þ q ghNR G 

    retðQNÞ þ P B:   ð88aÞ

    As with the dynamical equations derived for the driving phase, we see that the final dynamical equation in theretraction phase formally resembles a single lumped element, but with effective values for the various dynamicquantities that depend on the entire fluid path, starting from the reference port. We have,

    lNR   þ

     NS N

    S T ðlNT þ lT þ lTRÞ;   ð88bÞ

    and

    hNR   þ hNT þ hT þ hR:   ð88cÞ

    We will ignore the variations of  l N and  hN, with respect to the much larger quantities of  l NR and  hNR. In addi-tion, there is an effective ‘‘K ’’ constant for the retraction phase of 

     K retðQNÞ   K 

    toplossT   ðQNÞ  K 

    botlossT   ðQNÞ

    ðS T= N Þ2

      K lossN   ðQNÞ

    S 2N

      2

    S 2N

    !;   ð88dÞ

    and an effective ‘‘G ’’ quantity for the retraction phase,

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   161

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    22/29

    G retðQNÞ   G TðQNÞ

    S T= N   þ

    G NðQNÞ

    S N

    G R

    S N:   ð88eÞ

    With this last equation, we have completed the assembly of all the equations necessary to describe the dispenseprocess.

    5. Solution methodology

    The initial value problem of solder dispensing, defined by Eqs.  (83), (84), (86), and (88), were solved on aUNIX workstation with an implicit Adams–Moulton scheme available in the IMSL Library. Given materialproperties and geometric information pertaining to the pump, nozzles, diaphragm, and throttle, the modeltakes the drive pressure and diaphragm stop location as input and predicts the time dependence of followingquantities:

    1. The pressure fill of the pneumatic chamber upstream of the diaphragm.2. The diaphragm motion.3. The solder interface location, including the retraction phase.

    4. The solder interface velocity, including the retraction phase.5. The solder velocity in the throttle, including the retraction phase.6. The amount of solder dispensed.

    The numerical solution proceeded in two time intervals, that of the driving phase, and that of the retractionphase. Each phase has its own set of equations, as we have detailed above. We need to link these equationsacross the flow discontinuities caused by the diaphragm’s abrupt stop.  Appendix A  shows that the volumeflow rate in the nozzles just after the diaphragm hits the stop,  QþN, is related to the volume flow rate in thenozzles just before the diaphragm hits the stop,  QN, by the relation

    QþN  ¼ QN gQ

    D;   ðA:5Þ

    where Q

    D  is the diaphragm volume flow rate, just before the diaphragm hits the stop. The equation above al-lows us to determine the starting volume flow rate for the nozzle flow in the retraction, given the nozzle anddiaphragm flows just before the diaphragm hits the stop. In the equation above, there is a constant  g, which inprinciple depends on ratios of intrinsic length scales. However, in the numerical simulations that we per-formed, we adjusted the constant, experimentally.

    6. Results

    6.1. Parametric study

    We verified our dynamical model by comparing predictions from the model, applied to a prototype dis-penser, with dispensing data gathered from the same dispenser. Typical parameters pertaining to that dis-penser are given in  Table 1. The results of the study are summarized in  Figs. 6–12.

    Fig. 6 shows the calculated pressure transient in the pneumatic chamber, and the calculated diaphragm dis-placement transient through a single dispensing pulse. The stop is placed at a diaphragm displacement of approximately .04 mm (1.5 thousandths of an inch), and the diaphragm takes a little over 3 ms to reachthe stop.

    Fig. 7 shows the solder interface position in the nozzles as the pulse progresses, and Fig. 8 shows the veloc-ity transient of the fluid in the nozzles. The interface accelerates slowly at first (while the pressure in the pneu-matic chamber builds). Rapid acceleration follows until the diaphragm hits the stop. In the featured run, theinterface takes roughly 4 ms to reach the channel exit. A rapid change in velocity is experienced when the dia-phragm hits the stop. Since the throttle opening in the featured run is small, the nozzle flow is dominant overthe throttle flow just prior to impact. The opposing momenta of the two streams results in a significant reduc-

    tion in the nozzle flow, and a change in direction of the throttle flow, i.e. solder instantaneously begins to move

    162   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

    http://-/?-http://-/?-

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    23/29

    from the reservoir back into the body of the pump. Hence, the solder continues to dispense after the dia-phragm has stopped moving. However, the capillary forces in the non-wetting nozzles are strong enough toovercome gravity, and the momentum of the flow, and act to retard the flow. Eventually, the flow rate throughthe nozzles reaches zero, and subsequently becomes negative, i.e., the interface begins to retract. The retractinginterface achieves a maximum velocity of about 0.3 m/s before the interface reaches the top of the nozzles, atwhich point the capillary forces reduce dramatically (since the diameter of the pump body is much larger thanthat of the nozzles), and the interface motion ceases. The impact of throttle cross-section on the peak retrac-tion velocity is discussed later in this section.

    Fig. 9 shows the time dependence of the volume of solder that has passed through the nozzle exit. The vol-ume dispensed through the entire pulse, represented by the value at the end of the curve in  Fig. 9, is about

    1.25 mm3.

    Table 1Typical parametric values for model study

    Length of nozzle   l N   1.0 mmDiameter of nozzle   DN   80 lmNumber of nozzles   N    256Diameter of diaphragm   d D   20 mm

    Diaphragm ‘‘Spring Constant’’   k D   2.564 N/mDiaphragm stop distance   DxD   38.1 lmThrottle length   l T    8.0 mmThrottle gap width   R0 – Ri    130 lmCharacteristic length (diaphragm to throttle)   l DT   15 mmCharacteristic length (diaphragm to nozzle)   l DN   58 mmCharacteristic length (nozzle to throttle)   l NT   66 mmVertical height (nozzle to diaphragm)   hDNtop   51 mmSolder density   q   8000 kg/m3

    Solder viscosity   l   .002 kg/m/sSolder contact angle cos(h) 0.94Reservoir height   hR   15 mmFilling chamber temperature   T F   523 KGas pulse driving pressure   P G   4.116–5 Pa

       P  r  e  s  s  u  r  e   (   N   /  m   2   )

    1.00E+05

    2.00E+05

    3.00E+05

    4.00E+05

    5.00E+05

    6.00E+05

    7.00E+05

    8.00E+05

    0 1 2 3 4 5

    Time (ms)

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    1.6

    Pressure

    DiaphragmDisplacement

    Di   s pl   a c em en t  

     (  mi  l   s– t  h  o u s an d  t  h 

     s of   ani  n ch  )  

    Fig. 6. Pressure and diaphragm displacement transients.

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   163

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    24/29

    Fig. 10 shows the effect of widening the throttle opening. The upward fluid momentum in the pump due toflow through the throttle becomes significantly larger since the throttle presents a smaller resistance to the flowthan in the previous run. Thus, after the diaphragm hits the stop, dispensing ceases, and the solder interface inthe nozzles begins to retract rapidly. The velocity in the nozzles is actually negative immediately after the dia-phragm hits the stop. When the interface reaches the top of the nozzles, the fluid velocity is greater than 1 m/s.Hence, there is a greater propensity to draw gas into the pump from the nozzles if the throttle is opened wide.

    Fig. 11 shows the velocity transients of the flow through the throttle for two cases in which the solder levelsin the reservoir differ, in this case by 15 mm. To exercise the predictive capability of our formalism, we studiedhow the ‘‘quantity of solder dispensed’’ varies as the ‘‘height of solder in the reservoir’’ is changed. As the sol-der height in the reservoir increases, the upward flow through the throttle reduces (since a larger quantity of 

    solder must be moved), while the flow through the nozzles remains relatively unchanged. Due to the increase

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    0 2 4 6 8 10 12 14

    Time (ms)

       I  n   t  e  r   f  a  c  e   P  o  s   i   t   i  o  n   i  n   N  o

      z  z   l  e  s   (  m  m   ) Interface Position

    Fig. 7. Interface position transient of fluid in nozzles.

    -0.4

    -0.2

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    0 2 4 6 8 10 12 14

    Time (ms)

       V  e   l  o  c   i   t  y   i  n   N  o

      z  z   l  e  s   (  m   /  s   )

    Nozzle

    Fig. 8. Velocity transient of fluid in nozzles.

    164   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    25/29

    in mass in the reservoir, and thus an increase in the inertia of the solder, the diaphragm takes longer to hit thestop and, hence, the total quantity of solder dispensed increases.

    Fig. 12 shows the general trend of increasing dispensed solder volume as the solder height increases.

    6.2. Comparison of data and simulations

    To verify the fidelity of our dispensing model, experimental data was collected using the prototype appa-ratus described in Section 2. Measurements were extracted from high speed video of the dispensing operationlinked to the recorded diaphragm motion by means of a trigger, and the operating parameters were varied tocover a range of data. The model parameters, i.e., the characteristic length scales (the l ’s, particularly l DR, l DN,and   l NR), and  g, were determined by fitting the model predictions of 

    1. The dispensed solder volume.2. The time at which the interface reached the nozzle exit.

    3. The time at which the interface began receding.

    0 2 4 6 8 10 12 14

    Time (ms)

    Solder Volume

       V  o   l  u  m

      e   (  m  m   3   )

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    Fig. 9. Volume-dispensed transient.

    -1.50E+00

    -1.00E+00

    -5.00E-01

    0.00E+00

    5.00E-01

    1.00E+00

    0 0.5 1 1.5 2 2.5 3

    Time (ms)

       V  e

       l  o  c   i   t  y   i  n   N  o  z  z   l  e  s   (  m   /  s   )

    Nozzle Velocity (Wide Throttle)

    Fig. 10. Nozzle velocity for wide throttle.

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   165

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    26/29

    Figs. 13–15 show dispensed-solder transients for three different throttle openings ranging from ‘‘tight’’, orrelatively small, to ‘‘wide’’, or relatively large.

    Fig. 13 presents the results for the ‘‘intermediate throttle’’ case. The experimental displaced solder volume(1.17 mm3) compares favorably with the computed value (1.14 mm3). The time at which the interface reachesthe nozzle exit (3.2 ms), and the time at which the interface begins receding (5.8 ms) also compare well withthe computed values (3.2 and 5.7 ms respectively.) The comparison of the experimental and predicted values of these characteristic times is an excellent indicator of the ability of the model to capture the basic physics of thepump’s operation.

    Fig. 14 is a similar comparison for the ‘‘tight’’ throttle case. Fluid issues from the nozzle and disappearsfrom the field of view after about 4 ms, after which volume comparisons are not possible. Prior to this time,however, the volume displaced matches well with data. Also, the time at which the interface reaches the nozzleexit (3 ms) compares favorably with the computed value.

    Fig. 15  presents the results for the ‘‘wide’’ throttle case. The experimental data here appear somewhat

    scattered. However, the computed time taken for the interface to reach the channel exit is within 10% of 

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    1.6

    1.8

    90 95 100 105 110

    Static Head (mm)

       V  o   l  u  m  e

       D   i  s  p  e  n  s  e   d   (  m  m   3   )

    Solder Volume Dispensed

    Fig. 12. Solder volume dispensed vs. static head.

    -1

    -0.5

    0

    0.5

    1

    1.5

    2

    2.5

    3

    0 1 2 3 4 5 6

    Time (ms)

       T   h  r  o   t   t   l  e   V  e   l  o  c   i   t  y

       (  m   /  s   )

    Static Head = 90 mm

    Static Head = 110 mm

    Fig. 11. Impact of static head on throttle velocity.

    166   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    27/29

    the experimental value, the computed time at which the interface begins receding is within 20% of the exper-imental value, while the computed maximum solder volume displaced is within 30% of the experimentalvalue.

    We remark that, due to the difficulties associated with the dimension measurement of tiny drops from videophotographs, the accuracy of the experimental data is somewhat limited. Hence, the aforementioned compar-isons seem to indicate that the data are well represented by the model, and that the model can serve as a valu-able aid while determining the best operating conditions for solder dispensing. In particular, dispensing is mostconsistent when the solder flow stops abruptly immediately after the diaphragm hits the stop. If, instead, thesolder continues to ‘‘dribble’’ from the nozzles (as would happen if the throttle setting were too tight), occa-sional scatter of the droplets can result. Conversely, if the throttle is opened too wide, the peak retraction

    velocity of the interface may be too large and air may be drawn into the pump at the end of the pulse.

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    0 2 4 6 8 1210

    Time (ms)

       V  o   l  u  m  e   D   i  s  p  e  n  s

      e   d

       (  m  m   3   )

    Solder exits

    nozzle

    Solder retracts

    Model Simulation

    Experimental Data

    Throttle Area = 0.954 mm2

    Fig. 13. Experimental verification – intermediate throttle position.

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    0 2 4 6 8

       V  o   l  u  m  e   D   i  s

      p  e  n  s  e   d   (  m  m   3   )

    Model Simulation

    Experimental

    Throttle Area = 0.109 mm2

    Time (ms)

    Fig. 14. Experimental verification – tight throttle position.

    V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169   167

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    28/29

    7. Conclusions

    A mathematical model, employing fluid lumped elements, that describes molten solder dispensing from apneumatically actuated, diaphragm-driven pump was developed. The model is multi-physical in that capturesthe filling of the pneumatic drive chamber, the equation of motion of the diaphragm, and the flow of solder in

    the pump. The simulation capability enables predictive ‘‘tuning’’ of the pump. The model indicates that thesystem parameters should be chosen such that a desired amount of solder is first dispensed, followed soonafter by the diaphragm hitting the stop. The throttle opening should be selected to ensure that the solder flowis near zero immediately after the diaphragm hits the stop. The interface will then break cleanly, and theretracting solder proceeds back up the nozzles with a small velocity and acceleration, reducing the propensityto draw gas into the pump.

    Appendix A. Relation of flows across the discontinuity of diaphragm stop

    Here we display the continuity conditions of the flow variables at the point in time when the diaphragm hitsthe diaphragm stop. At this time there is a step discontinuity in the flow driven by the diaphragm resulting in

    step changes in the flow at the other two ports. Let us focus again on the main chamber, which has three ports,the diaphragm port D, the bottom throttle port T, and the Nozzle ports N. If one integrates Eq.  (11),

    qX N 

    q

    d

    dt ð M  pqQqÞ þ

    1

    2q

     K  p 

    S 2 p Q2 p  þ q gZ  p  þ P  p  ¼ 0; p  ¼ 1; . . . ; N    ðA:1Þ

    over a vanishing small time interval around the time of the diaphragm impact, the only terms that survive are:

    Port D: M DDDQD þ M DTDQT þ M DNDQN þ

    Z   ee

    dtP D=q ¼ 0;   ðA:2aÞ

    Port T: M TDDQD þ M TTDQT þ M TNDQN  ¼ 0;   and   ðA:2bÞ

    Port N: M NDDQD þ M NTDQT þ M NNDQN  ¼ 0:   ðA:3cÞ

    0.00E+00

    2.00E-01

    4.00E-01

    6.00E-01

    8.00E-01

    1.00E+00

    1.20E+00

    0.00E+00 2.00E+00 4.00E+00 6.00E+00 8.00E+00

    Time (ms)

    Model Simulation

    Experimental Data

    Throttle Area = 1.35 mm2

       V  o   l  u  m  e   D   i  s  p  e  n  s  e   d   (  m  m   3   )

    Fig. 15. Experimental verification – wide throttle position.

    168   V. Jairazbhoy, R.C. Stevenson / Applied Mathematical Modelling 32 (2008) 141–169

  • 8/19/2019 1-s2.0-S0307904X06002927-main

    29/29

    Here, DQ p is the change in flow rate across any flow discontinuity at por