Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Applied Thermal Engineering 31 (2011) 656e667 Contents lists available at ScienceDirect Applied Thermal Engineering journal homepage: www.elsevier.com/locate/apthermeng Prediction and evolution of drop-size distribution for a new ultrasonic atomizer Moussa Tembely a, *, Christian Lecot b, Arthur Soucemarianadin a a b University of Grenoble, Laboratory of Geophysical and Industrial Fluid Flows, UMR 5519, BP 53, 38041 Grenoble, France University of Savoie, Laboratory of Applied Mathematics, UMR 5127 CNRS, 73376 Le Bourget-du-Lac, France a r t i c l e i n f o a b s t r a c t Article history: Received 14 December 2009 Accepted 28 September 2010 Available online 7 October 2010 Complete modeling of a new ultrasonic atomizer, the Spray On Demand (SOD) printhead, was carried out to enable its optimization. The modeling was focused on various factors, including nozzle vibrations and a theoretical prediction of the SOD drop-size distribution. Assuming that the spray is generated based on Faraday instability, a prediction of the drop-size distribution within the framework of a specific and general Maximum Entropy Formalism (MEF) was developed. This prediction was formulated using the conservation laws of energy and mass, as well as the three-parameter generalized Gamma distribution. After establishing an analytical expression to estimate the Sauter Mean Diameter, a qualitative validation of the model was performed by comparing predictions with experimental measurements of the dropsize distribution. The dynamic model is shown to be sensitive to operating conditions and physical properties of the fluid. The prediction capabilities of the model were found to be adequate, paving the way for optimization of the atomizer. The evolution of the drop-size distribution, under the coalescence effect, was also assessed using a convergent Monte Carlo method to solve the distribution equation. This was formulated in a mass flow algorithm, leading to a more physically relevant distribution.  2010 Elsevier Ltd. All rights reserved. Keywords: Nozzle conveying fluid Spray Maximum entropy formalism Monte Carlo method Coalescence 1. Introduction It is well known that piezoelectric jet printing technology is capable of depositing controlled amounts of fluid onto a specified location very accurately. Combining this ability with an increasingly wide selection of fluids has made the piezoelectric on-demand jetting system a promising device for the development of innovative industrial applications. Jet printing technology offers an amazingly broad utilization range across a wide variety of applications, such as biotechnology, electronics, pharmacology, micro-optics, and many others that are only limited by our imagination [1]. Four major fluid jetting techniques are commonly used: Drop On Demand (DOD), Continuous Ink Jet (CIJ), Electro-Valve and Spray Technologies. The work presented here focuses on an innovative spray technology that may be qualified as a Spray On Demand (SOD) technology, where spray is generated only when required. This constitutes a major difference with classical spraying technologies, in which the jetting is continuous. Inkjet printheads are advantageous because they are non-contacting, minimizing the potential for contamination. In general, classical printheads are well adapted to applications that deal with Newtonian fluids. However, a growing number of * Corresponding author. Tel.: þ33 456 521 120; fax: þ33 475 561 620. E-mail address: moussa.tembely@ujf-grenoble.fr (M. Tembely). 1359-4311/$ e see front matter  2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.applthermaleng.2010.09.027 engineering applications need robust devices that allow for the ejection of rheologically complex fluids (e.g., polymers and fluids with particles). In this paper, we present the underlying theory for the optimization of an innovative Spray On Demand printhead (SOD) that is under development. This printhead has the advantage of being robust and relatively simple, enabling high-throughput printing. In contrast to other printing techniques like CIJ or DOD, this new device is able to eject fluids that are much more viscous or that contain particles, for example, while lowering the risk of clogging. As shown in Fig. 1, the SOD consists of a tube or a nozzle that conveys fluid. A piezoceramic disk is welded at the middle region of the tube. The tube motion is induced by vibration of the piezoceramic element, leading to spray generation via a hydrodynamic instability. Relating the tube motion to the resulting droplet-size distribution is of paramount interest for device optimization, especially considering that the device has already been used for gold electrode manufacturing on fabrics and is being considered for use in fuel cell manufacturing [2]. Thus, atomization constitutes the main focus of this work. Despite the large number of potential applications for this technology, the modeling of atomization and droplet formation remains an open problem. This effort is inherently problematic [3e5] and requires extensive efforts to collect accurate data and models with a significantly broad range of validity. A general theory to predict the drop-size distribution of a spray remains an unsolved problem, although some distributions (e.g., M. Tembely et al. / Applied Thermal Engineering 31 (2011) 656e667 657 Fig. 1. Spray On Demand printhead and visualization of the spray. Weibull and Rosin-Rammler distributions) follow from the physically-grounded percolation model of chaotic atomization [6]. In the end, the application of these theories ultimately leads to curve fitting. The generated droplets result from the dynamics of ligaments and their fragmentation [7]. An analysis of ligament fragmentation based on the assumption that it consists of sub-drops in different sub-layers has been performed [8]. In this case, the ligament breakup was well represented by gamma distributions. All of these models ultimately lead to distributions that still must be fit with experimental data and do not account for the effect of the atomizer. The significant number of drops that constitute a spray does not allow for the precise determination of the diameter or velocity of each drop. There is large body of literature that discusses the prediction of a representative drop diameter in a spray. However, there are relatively few publications dealing with the prediction of drop-size distributions. One possibility to quantitatively describe a spray is to adopt the tools of statistical analysis. Following a previous report [9], there are three methods for modeling the drop-size distribution: Empirical Method, Probability Function Method and Maximum Entropy Formalism (MEF), which is the one adopted in the present work. Despite its large presence in industrial applications, spray modeling remains a challenge for computational methods and experimental measurements when one wants to predict the dropsize distribution. Droplet generation is an extremely complex process that cannot be precisely determined. Current approaches are either semi-empirical or need to be adjusted to each operating condition encountered. Based on ultrasonic atomization through the Faraday instability of the SOD, we propose a drop-size distribution that is based on physical evidence and is sensitive to operating conditions. This type of approach is necessary to obtain a specific drop-size distribution, which may be required in some applications. Extending the MEF drop-size distribution to enable the generation of a temporal evolution was also investigated. This could also be of great interest for initializing sophisticated CFD spray modeling codes. In traditional MEF approaches, the precise drop-size distribution is usually timeindependent. Here we propose a modeling of the time-dependent drop-size distribution of the MEF. First, the distribution evolution equation is derived and the Mass Flow Algorithm (MFA) is utilized to simulate the equation with a convergent Monte Carlo method. 2. Spray generation with a vibrating nozzle The SOD nozzle that conveys fluid is excited to an ultrasonic mode at a frequency of fp ¼ 192 kHz using a piezoceramic actuator. Subsequently, the pendent drop or film at the beveled nozzle tip breaks up into droplets via a Faraday instability, establishing that the frequency of the stationary surface wave corresponds to half of the imposed working frequency (e.g., by the SOD nozzle) [10]. Based on experimental investigations [11,12], the mean droplet diameter (D30) is proportional to the surface wavelength (l) for ultrasonic spray generators, favoring capillary wave theory [13]. The crests of these surface waves breakup to generate droplets. The droplet volume can be seen as a fraction of the crest volume generated by the intersection of unstable standing waves. Equating the mean droplet volume with a fraction of the crest volume leads to the following relation: 2 D330 fl A (1) where A is the amplitude of the film at which an instability leads to droplet ejection. Using the fact that near the breakup point D30 fl, the following relation can be deduced from (1): (2) Afl To predict the droplet size generated from the surface wave, nonlinear dynamics analyses of the wave dispersion and interactions should be performed. The few non-linear analyses available [14] show that a square wave structure is formed preferentially for small viscous liquids at frequencies >100 Hz. For this purpose, an empirical relationship has been established in [15] for the unstable wavelength that relates all of the main parameters, including the fluid viscosity mf, for ultrasonic atomization (100 kHz) of a pistonlike inducement of the Faraday wave. This relationship can be expressed as follows: 0 11=5 mf sf A r2f fp3 lz@ (3) where rf and sf are the density and the surface tension of the fluid, respectively. A quick analysis of Eq. (3) can be carried out by turning it into a dimensionless form and by assuming that the unstable wavelength is expressed as l ¼ Pðs; m; fp ; rÞ. Taking the wavelength as a characteristic length (we could also choose the wave amplitude A instead knowing Afl), the velocity vc ¼ lfp and the previous relationship given in Eq. (3) is simply equivalent to Re:We ¼ Const, where Re ¼ rf vc l=mf and We ¼ rf v2c l=sf . Thus, this analysis relates the unstable wavelength to all of the main forces involved: inertia, viscosity and the surface tension effect. We prefer this relation to the one given by Lang [11], which does not account for the viscosity. 3. New physically based MEF Following [7], three interpretations have been attributed to the fragmentation process: (i) a sequential cascade of breakups, which is an interpretation that originated from Kolmogorov and leads to the log-normal distribution; (ii) the aggregation scenario, which makes use of the Smoluchowski kinetic aggregation process and leads to the fact that the drop-size distribution displays an 658 M. Tembely et al. / Applied Thermal Engineering 31 (2011) 656e667 exponential tail; and (iii) the Maximum Entropy Formalism (MEF), which leads to a Poisson distribution, as presented by the author. It is worth adding to these interpretations the idea of the dynamics of ligaments, which assumes that they are constructed of blobs of different layers [8]. Instead of being opposed, we show that these different interpretations could be complementary to some extent. From two visions of the MEF, we derive a new formulation for predicting drop-size distribution. This work shows that the MEF could lead to the same result given by ligament dynamics, which are at the core of spray formation. The droplets coming from the breakup of these different ligaments can explain the polydispersity of the spray, even though ultrasonic sprays are less disperse compared to other techniques [16]. To establish a physically based approach, we propose a coupling of what we distinguish as two complementary formulations using MEF, which we designate as specific and general formulations. A specific formulation is based on conservation laws; this formulation is proven to give satisfactory results on the spray, with a drawback existing for small-size drops. Its advantage is that it takes into account the device operating parameters. Thus, we propose to couple this approach with a general formulation, leading to the three-parameter generalized gamma distribution. Indeed, this distribution takes into account the general characteristics of a spray, but does not take into account the unique features of the atomization device. the spray. In [17], a distribution is derived based only on the constraint of mass conservation, and the deduced volume-based distribution gives good results for a pressure swirl atomizer. The authors applied the MEF not on the number-based fraction, but on the volume-based fraction (which is not a probability, as explained in [18]). They end the method by performing a variable change to deduce the number-based drop-size distribution. Other attempts have been made by exploiting this variable change to end up with an acceptable number-based distribution. In fact, the challenge of this specific formulation is to correctly predict the number-based distribution especially for small droplet populations, which are mainly over predicted. To prevent this over-prediction, the “variable change” has been carried out, but it violates the MEF principle [18]. Since the introduction of the general approach [19], the previous formulation has somehow been abandoned. The approach presented here revisits the abandoned specific formulation and shows its complementarity with the general formulation approach, as well as its consistency with the MEF principle. In fact, it is unclear how one can claim to predict a process that is as complicated as the atomization without using as much information as possible, including the useful information given by the specific formulation. Many aspects have an influence on the generation mechanism of the droplets, the atomizer itself and the fluid. In addition to these constraints, a practical approach for spray modeling is to use the minimum number of parameters to enable effective computation. Our approach aims to meet these requirements by coupling the presented specific formulation and the following general formulation. 3.1. Specific formulation 3.2. General formulation This formulation, which is sensitive to SOD operating conditions and based on spray conservation laws, consists of maximizing the following Shannon entropy: S ¼ Z (4) hn ðDÞlnhn ðDÞdD D where hn(D) is the number-based drop-size distribution of the spray. The integration is performed on all of the permissible states R Dmax/N R , where Dmin, Dmax are the spray drop of the diameter h Dmin/0 D minimum and maximum diameters, respectively. The associated constraints to Eq. (4) are as follows: Normalization constraint Z hn ðDÞdD ¼ 1 (5) D Conservation law constraints Z hn ðDÞmk ðDÞdD ¼ Hk ; k ¼ 1.M (6) As previously mentioned, all atomization processes can be seen as a breakup of a ligament. Therefore, ligament dynamics can be seen as a general characteristic of a spray. Because of the random liquid motions present in a ligament, the sub-drops of a layer overlap and merge. The solution of that evolution equation [8] leads to the following gamma function or gamma number-based distribution: FLn ðx ¼ D=d0 Þ ¼ bb b 1 x exp½ bxŠ GðbÞ (8) where d0 is the average blob or ligament diameter; and b is the gamma distribution parameter. Consider that the ligament population at the SOD nozzle is distributed as follows: FL ðd0 Þzdðd0 (9) dl Þ Unlike the case for the high speed jet ligament, the ligament from the Faraday wave is less disperse, and assumed to have nearly the same dimension, noted here as dl. Therefore, we can deduce the drop-size distribution of the spray (FSn(D)) using convolution of FLn(D/d0) and FL(d0) as follows: D where mk (D) is a function describing the state of the droplets (e.g., volume or energy); and Hk is the available information and known moments of the distribution. The solution of the system given in Eqs. (4)e(6) can be expressed in the continuous form: c hn ðDÞ ¼ exp l0 X k c ! lck mk ðDÞ FSn ðDÞ ¼ where lk are the Lagrange multipliers in continuous form. Until recently, this formulation of the MEF was applied using the conservation laws of mass, momentum or energy. Its predictive capability was quite satisfactory for the volume-based distribution of dðd0 Þ 1 FL ðd0 ÞFLn ðD=d0 Þ ¼ FLn ðD=dl Þ dl dl (10) We can finally deduce the following: b FSn ðDÞ ¼ (7) Z 1 b b ðD=dl Þ dl GðbÞ 1 exp½ bðD=dl ފ (11) This gamma distribution could therefore represent the general description of a spray (e.g., that of the SOD spray). Adoption of the MEF formalism can also produce an alternative description of a spray that leads to the gamma function. This distribution is formulated with a single constraint on the diameter expressing the definition of a mean drop diameter [19]: 659 M. Tembely et al. / Applied Thermal Engineering 31 (2011) 656e667 Z Dq Fn ðDÞdD ¼ Dqq0 (12) D where and q and Dq0 are two parameters of the distribution function, Fn(D). To model the small drop limitation caused by the presence of surface tension forces, a diameter-class probability distribution, g(D), that continuously increases with the diameter is introduced [20]: 1 gðDÞ ¼ 2Da where 2 is a constant and a is the third parameter of Fn(D). This yields the three-parameter generalized distribution [20,21]: "  a a 1 a qD q Fn ðDÞ ¼   exp a G a q Dq0 q a D q Dq0 !q # Z Fn ðDÞln Fn ðDÞ dD mðDÞ (14) (15) D where mðDÞ ¼ gðDÞa 1 , which is the probability of reaching a diameter D. The specific approach enables determination of some parameters of the general formulation, as follows: Dq0 ¼ 8 <Z : Hk ¼ Z D 91=q = Dq hn ðDÞdD ; mk ðDÞFn ðDÞdD; k ¼ 1.M We can express Ms differently in the following manner: Ms ¼ rf Vspray ¼ This is carried out by maximizing the following relative entropy: SR ¼ Fig. 2. Drop hanging at the nozzle tip and atomization modeling. (13) (16) (17) D The constraints should be chosen to determine the three parameters (Dq0,a,q) of the generalized gamma function, in addition to Eq. (16). These constraints should be as relevant as possible and should reflect the physics of atomization. Note that the specific formulation that we introduced was solely able to predict the dropsize distribution, although it poorly models the distribution for small drops [9,18]. 4. Application to the SOD spray Predicting the drop-size distribution of the new Spray On Demand (SOD) printhead is an important issue for device optimization and operation. These steps are required to further extend device operation. In the following section, we apply the previous approach to the SOD with the details of the specific formulation constraints on mass and energy conservation. 4.1. Mass conservation Let Ms be the mass of fluid accumulated at the nozzle tip. This mass is ejected after an excitation time or operation of the SOD and will breakup into droplets. It corresponds to the total mass of fluid bursting at the nozzle tip interface due to Faraday instabilities, which is a direct consequence of the acceleration. The mass Ms of the fluid is a reference quantity on which we will make our argument of mass and energy conservation. Therefore, its precise value is not significant. We assume Ms to be atomized into spray droplets with the drop diameter space divided into nc classes of constant width DD. nc X mi ¼ i¼1 nc X rf Vi ¼ rf i¼1 nc X (18) rf Vi i¼1 nc pX nc p X p Ni D3i ¼ rf N p D3 ¼ rf ND330 6 i¼1 6 i¼1 i i 6 (19) where pi is defined as pi ¼ Ni/N and D30 is the volume mean diameter. From Eq. (19), we can deduce the constraint upon mass conservation as follows: nc X pi d3i ¼ 1 with di ¼ i¼1 Di D30 (20) Before starting with the second constraint, the parameter Ms can be expressed in terms of operating conditions and fluid properties. For this purpose, a global analysis is carried out to determine Ms. As suggested by experimental observation, atomization takes place preferentially in the second half of the nozzle tip exit region (Fig. 1). The surface tension is defined as the force per unit length acting tangentially to the liquid-gas surface. By applying Newton’s second law to a drop hanging at the tip, including a balance between surface tension and gravity, we can deduce the following when projecting on a vertical component (Fig. 2): Ms gzFsf cosðp=2 ðqE ab ÞÞ ¼ sf CsinðqE ab Þ (21) where the force Fsf ¼ sf C, C is the boundary of the hanging drop. The value of Ms corresponds to the fluid mass (drop) outside the nozzle, which breaks up into spray; and ab represents the beveled angle of the nozzle tip. In Eq. (21), qE is the equilibrium contact angle between the liquid fluid and the nozzle structure. This value is set based on the back pressure so that qE > ab. The parameter Ph is the circumference of the elliptical nozzle exit shape and is expressed as follows: Ph ¼ 4aEðeÞ (22) The circumference of an ellipse is Ph ¼ 4aE(e), where the function E(e), with e being the eccentricity, is the complete elliptic integral of the second kind. A good approximation from Ramanujan is as follows:  Ph zp 3ða þ bÞ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð3a þ bÞða þ 3bÞ (23) where a is the semi-major axis and b is the semi-minor axis (Fig. 2). Also, using the beveled angle ab we obtain a ¼ ri/sin(ab) and b ¼ ri, where ri is the inner tube radius. Finally, C is expressed as follows: C ¼ 2ri þ  p Ph ¼ 2ri þ 3ða þ bÞ 2 2 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð3a þ bÞða þ 3bÞ (24) 660 M. Tembely et al. / Applied Thermal Engineering 31 (2011) 656e667 Then, Eq. (32) can be rewritten in the following form: 4.2. Energy conservation The instability leading to droplet formation can be viewed as the conversion of the surface energy, Esurface, and fluid film kinetic energy, Evib, of the pendent drop (of mass Ms) at the nozzle exit to the surface energy of the droplets, Edroplets, generated, in addition to the dissipation due to fluid viscosity: (25) Esurface þ Evib zEdroplets where we have neglected the other energy forms of the spray droplets (i.e., kinetic energy due to the low velocity of the ultrasonic spray [16]), gravitational potential and dissipation. Expressing the different terms of Eq. (25), we approximate the following: Esurface zsSs ¼ sf nc X 2 i¼1 Finally, the system for the MEF to be solved could be summarized as follows: nc X SD ¼ pab (26) 2 1 Evib z Ms < Vf2 >0;Tp 2 (27) where Tp ¼ 2p/up,up or fp are the piezoceramic disk applied pulsation and frequency, respectively. The displacement of a point on the nozzle tip drop interface can be of the following form: u  p t þ Bsin up t 2 (28) where the first term is the displacement due to the Faraday wave, which is vibrating at half the tube tip pulsation up; and the second term B is the nozzle tip displacement, which is sensitive to nozzle and piezoceramic disk proprieties, as expressed in Ref. [2] from the SOD nozzle motion analysis. Thus, the velocity can be expressed as follows:   A Vf ðtÞ ¼ up cos up =2 t þ Bcos up t 2 (29) RQ Calculating the mean value using hi0;Q ¼ ð1=QÞ 0 dt, we can finally deduce the following: 1 ~2 Evib z Ms < Vf2 >0;Tp ¼ p2 Ms fp2 A 2 (30) ~ 2 ¼ ðA=2Þ2 þ B2 . where we set A For the spray droplet energy, we can obtain the following relation: Edroplet ¼ sf Si ¼ psf N i¼1 nc X pi D2i (31) i¼1 Substituting Eqs. (26), (30) and (31) into (25), we obtain the following: 1 2 ~2 nc X 1 @ab pMs fp A 2 A pi Di ¼ þ sf N 2 i¼1 0 (32) From Eq. (18), we have the following expression: N ¼ Ms rf p6 D330 ¼ pi lnðpi Þ (35) i¼1 nc X pi ¼ 1 nc X (36) pi d3i ¼ 1 with di ¼ i¼1 nc X pi d2i ¼ i¼1 nc X (34) i¼1 where Ss is half of the elliptic surface of the nozzle tip and is taken here as an approximation of the hanging drop surface. Considering that the liquid film oscillation at the nozzle tip is the result of the tube vibration and capillary wave, the mean fluid film kinetic energy can be approximated as follows: df ðtÞ ¼ Asin 13 2 pfp2 A~ abg A5 þ sf 6 2sf sinðqE ab ÞC 0 p ¼ D30 4rf @ pi d2i sf sinðqE g ab ÞC 1 rf p6 D330 (37) D30 ¼ kc Dc (38) where Dc corresponds to a characteristic diameter of the process equivalent to the Sauter Mean Diameter D32 for the energy constraint, 0 13 2 pfp2 A~ p@ abg 4 A5 þ Dc ¼ 1= rf sf 6 2sf sinðqE ab ÞC 2 In fact, we can rewrite the energy constraint in the following form: nc X pi D2i Pnc i¼1 ¼ i¼1 pi D2 $ Pnci Pnc i¼1 p D3 i¼1 i i pi D3i ¼ D330 D32 (39) and can deduce by identification that Dc h D32. At the moment of droplet ejection, we assume that the amplitude is proportional to the wavelength. Then we take the following expression, using Eqs. (2) and (3): 11=5 0 mf sf A r2f fp3 Az@ (40) Finally, we can establish a relationship for estimating the Sauter Mean Diameter of the SOD, depending on the physical, mechanical and operating conditions of the printhead: D32 2 0 12=5 313 pfp2 1 mf sf abg 4 @ A þB2 5A5 þ sf 4 r2f fp3 6 2sf sinðqE ab ÞC 2 0 p ¼ 4rf @ 1 (41) A numerical application of this relationship for typical values of the SOD operation in [2], with B ¼ 5.06 mm for fp ¼ 192 kHz, leads to a Sauter Mean Diameter of 20 mm, which is in good agreement with the experimental measurement of 21 mm. Solving the MEF system leads to the following expression: pi ¼ exp (33) Di D30  l0 l1 d2i l2 d3i  (42) Using the result from [22], the following minimization enables determination of the Lagrange multipliers: M. Tembely et al. / Applied Thermal Engineering 31 (2011) 656e667 ( " nc X min ln exp i¼1   l1 d2i kc    1 l2 d3i #) gives l1 ; l2 (43) The multiplier l0 is given from Eq. (36) by the following relation: l0 ¼ ln " nc X i¼1  exp l1 d2i l2 d3i  # (44) Hence, the number-based drop-size distribution of the specific formulation can be expressed as follows: pi hn ðDi Þ ¼ (45) DD Knowing hn(D), the volume-based distribution fv(D) is deduced using the following relationship: fv ðDÞ ¼  D D30 3 (46) hn ðDÞ Using Eq. (3), we can determine D30 with a proportional constant 1 as follows: D30 11=5 0 mf sf ¼ z1 @ 2 3 A rf fp (47) where z1 could be determined using experimental measurement. 5. Predicting a physically based drop-size distribution The specific approach is limited by the fact that hn(D) generally over-predicts the population of small droplets [9]. The generalized gamma function, which was found to be identical to a NukiyamaTanasawa distribution [18,21], is in good agreement with most experimental measurements both for number-based, fn(D), and volume-based, fv(D), distributions [23]. Thus, for the modeling used in our approach, we derived the general formulation using the three-parameter generalized gamma distribution: Fn ðDÞ ¼ "  a a 1 a qD q   exp a G a q Dq0 q a D q Dq0 !q # (48) For the proposed method, we used the physically based approach of the specific formulation to compute the diameter Dq0 and parameter , respectively, using the following relationship: Z Dq0 ¼ 8 <Z : D D3 Fn ðDÞdD 91=q = D Dq hn ðDÞdD and D32 ¼ Z ; D2 Fn ðDÞdD allows our model to be sensitive to the physical and operating conditions of the device. It is worth mentioning that MEF is traditionally applied to a spray by determining the different parameters of a drop-size distribution function. Mathematically, this could be achieved using any combination of independent characteristics of the distribution (e.g., mean drop diameters). However, the resulting function is highly dependent on the characteristics utilized, and the desired solution is usually obtained for a unique set of information (data). The specific formulation allows the model to assure the uniqueness of the resulting distribution consistent with the MEF principle [24]. This can be performed without the need to change any variables, as is traditionally done by applying MEF to the spray. In other words, this invalidates the choice of some particular moments to infer the drop-size distribution. 5.1. Experimental measurements 5.1.1. Shadow imaging technique setup The spray was illuminated in a backlit configuration using a non-coherent short flash source (15 ns in duration) to freeze the droplet movements on the images. The detector of a monochrome camera was comprised of 1008  1018 square pixels that were 9 mm on each side. An objective composed of two lenses (foc1 ¼ 300 mm and foc2 ¼ 100 mm) and with a lateral magnification of G ¼ 3.7 was used to obtain high-resolution images. The field of view was 2.44 mm  2.46 mm and the resolution was 4130 pixels/cm. A three-step image-processing program developed in Cþþ was applied to the images [25]. The first step consisted of the normalization of the images to enhance their contrast and to correct for non-uniform background illumination. In a second step, the images were binarized using two thresholding techniques. The following thresholds were applied: a classical threshold based on the range of grey levels in the image and a threshold based on the wavelet transform (to detect the out-of-focus droplets). This technique enables the detection of local grey level variations for low contrast images to localize the maximum number of droplets and to associate a surrounding mask to each of them. Subsequently, each droplet was separated from its neighbors within the mask to enable their individual analysis. The sub-pixel contour of the droplet was then computed. The diameter of each droplet is defined as the equivalent surface radius r61 of the binarized image of the droplet at a level equal to 61% of the local grey level amplitude. An imaging model has been developed to correctly estimate the diameter [26]. The ratio of the real diameter D to the equivalent surface radius r61 is directly related to the contrast (C0) of the droplets, as shown in Fig. 3. This curve was used to estimate the real diameter of the (49) D where hn(D) is given using Eq. (45). We can show the following expression using Eq. (48): D32 ¼ D330 D220 1 ¼ 1  a q Dq0 qq G  G 2þa q 3þa q   (50) In this case, the parameter q is fixed using experimental validation; this parameter is mainly sensitive to the atomization process [23] and has a unique value for a given atomizer. Therefore, it is assumed to be constant for our model. This approach allows for a dynamic drop-size distribution. This is in contrast to the classical approach, where the parameters must be adjusted for each operating condition. The coupling of the previous physical method 661 Fig. 3. Calibration curve, droplet diameter evaluation. 662 M. Tembely et al. / Applied Thermal Engineering 31 (2011) 656e667 Fig. 4. Theoretical and experimental drop-size distributions of a new Spray On Demand printhead. drops. Moreover, the out-of-focus level of each droplet was determined to sort the droplets relative to their spatial position with respect to the focus plane. This was achieved through the calibration of the Point Spread Function (PSF) of the optical setup. Indeed, the PSF gives information on the out-of-focus level of the droplets [26]. Droplets with a low contrast level (<0.1) and high out-of-focus level (PSF >0.1 mm) were thus rejected. 5.1.2. Model constants determination We experimentally determined the drop-size distribution of the new printhead using a high speed camera and image post-processing treatment (shadow imaging technique). An adjustment was made to determine the two constants of the model. Once these parameters were determined, the model was able to predict the behavior of the device when physical and operating conditions were changed. It should be noted that the model takes into account the different physical parameters (e.g., fluid and nozzle proprieties), as well as the many other parameters required for SOD operation. From comparison with experimental results (Fig. 4), we deduced the two constants of our model by non-linear regression to the experimental data based on the least-squares method (Table 1), with a coefficient of determination R2 ¼ 0.9887, which is an excellent result. Fig. 5. Prediction of the physical model drop-size distribution. As an illustration of the model prediction capabilities, three operating vibration frequencies were applied to the PZA, and the results are given in Fig. 6. These different results are in good agreement with experimental results [27] for an ultrasonic atomizer. In Fig. 7, we present the drop-size distributions for different equilibrium contact angles of the fluid/nozzle. The effect of this parameter was found to be negligible. These types of predictive analyses are out of range for the traditional MEF approach. Other prediction capabilities can be found in Ref. [2], including the sensitivities to the SOD nozzle length or the Young Modulus. These conception parameters could potentially be chosen to produce the desired drop-size distribution, thanks to the proposed physically-based model. 6. Drop-size distribution evolution equation In previous sections consideration, the distributions have been time-independent. We assume that the real physical drop-size distribution varies with time (e.g., under the effect of coalescence and breakup). In the following section, we undertake an analysis that combines the physically-based MEF approach with the balance population method, allowing the evolution of the drop-size distribution. The evolution of the distribution function is governed by a Boltzmann-type equation [28]. We focus on the effect of 5.2. Drop size distribution: model prediction capability Our main goal was to test the ability of the model to qualitatively predict the drop-size distribution to yield tools for the optimization of the SOD. For a quantitative study, further measurements would be necessary and would require a reliable SOD. The results reported in Fig. 5 are predictions of the model with respect to various physical properties of the fluid. The curves presented here show that the drop-size distribution becomes more dispersed with increasing surface tension, the most relevant parameter. A decrease in the surface tension leads to a finer spray of droplets, where the distribution shifts toward the small droplet population with an increase in the main peak height. This trend is in good agreement with capillary instability, assuming the unstable wavelength is related to the spray droplet size (see Eq. (2) or [11]). Table 1 The constants of the model. z1 q 2.98. 0.21 Fig. 6. Volume-based drop-size distribution prediction sensitivity with respect to the piezoceramic frequency. 663 M. Tembely et al. / Applied Thermal Engineering 31 (2011) 656e667 i vnði; tÞ 1X kc ði ¼ vt 2 j¼1 j; jÞnði j; tÞnðj; tÞ nði; tÞ nc X kc ði; jÞnðj; tÞ j¼1 (56) where we set kc ði; jÞ ¼ Kc ðVi ; Vj Þ and nðVi ; tÞ ¼ nði; tÞ. 6.1. Coalescence kernel determination One difficulty in the present approach is to correctly write down the kernel that expresses the coalescence of drops. We express the coalescence kernel as the product of the coalescence efficiency (Le) and the collision frequency (Hf): Kc ðV; VÞ ¼ Ka Le V; V 0 Hf V; V 0 (57) where we have introduced Ka as an adjustable constant that depends on the situation. Fig. 7. Volume-based drop-size distribution prediction sensitivity with respect to the equilibrium contact angle. coalescence and neglect the breakup, evaporation and nucleation phenomena. Using the drop-size distribution in a formulation where it depends only on time t and volume V (or diameter D), the balance population equation for the distribution fn(V,t) can be expressed as follows: v½NðtÞfn ðV;tފ 1 ¼ vt 2 ZV Kc V V 0 ;V 0 NðtÞfn V NðtÞfn ðV;tÞ Kc V;V 0 NðtÞfn V 0 ;t dV 0 ð51Þ where N(t) is the total number of particles at time t. The right-hand side (rhs) term represents the source and sink effects due to coalescence. In this work, we consider only binary interactions where broken droplets split into two smaller ones and where two droplets can coalesce to form a bigger one (before impacting on the substrate). The probability of finding a drop with a volume between Vi and Vi þ DV is the same as the probability of finding a drop with a diameter between Di and Di þ DD. The change between the volume and diameter formulations is carried out using the following relationship: fn ðDÞdD ¼ fn ðVÞdV (52) Then, we can deduce the following relation: 2 pD2 fn ðDÞ (53) where fn(V,t) is the number-based drop-volume distribution to be determined by our analysis. The equation relating the number fn(D) and volume fv(D) based drop-size distributions is given by the following equation: fv ðDÞ ¼  D D30 3 fn ðDÞ (54) The total number of droplets in the range V þ dV at time t is expressed as follows: nðV; tÞ ¼ NðtÞfn ðV; tÞdV tcoal V; V 0 =tcont V; V 0 (55) Then, we can deduce the discrete form of the equation that is verified by the average number of droplets in class i, n(Vi, t) as follows: (58) where tcoal ðV; V 0 Þ; tcont ðV; V 0 Þ are, respectively, the average coalescence time and the contact time of particles with volumes V and V 0 . The time required for coalescence can be estimated [30] using the following relation: tcoal V; V Vmin fn ðVÞ ¼ Le V; V 0 ¼ exp V 0 ;t NðtÞfn V 0 ;t dV 0 Vmin V Zmax 6.1.1. Coalescence efficiency Following [29], we assume that the coalescence efficiency could be expressed as follows: 0 ¼ C1 R3V;V 0 rf sf !1=2 (59) where C1 is a constant close to 2. The parameter RV;V0 is the equivalent radius of the radii of coalescing drops and is defined as follows: RV;V 0 ¼  1 1 þ DðVÞ D V 0  1 (60) where DðVÞ ¼ ð6V=pÞ1=3 . The contact time is estimated in Ref. [31] for a flowing fluid, including the contribution due to relative velocities between bubbles and is assumed here for droplets: tcont V; V 0 ¼ DðVÞ þ D V 0 2ur V; V 0 (61) Here we have neglected the turbulence effect. We denote ! u r ðV; V 0 Þ as the relative velocity between drops of volumes V and V 0 . The determination of this relative velocity is performed as follows. The relative velocity can be expressed by estimating the terminal velocity of falling particles. We assume in our model that coalescence happens only after this regime is reached. This is reasonable because the momentum velocity response time ðzD2 ra =ma Þ is small, considering the micro-size scale of the spray droplets. From Newton’s second law of falling particles, once the terminal velocity is reached we have the following relation: R2ea CD ¼ 4 Ga 3 (62) where Rea ¼ ra Vl D=ma and Ga ¼ D3 gðrf ra Þra =m2a are, respectively, the Reynolds and Galileo (or Archimedes) numbers; rf, mf, (ra ,ma) are, respectively, the density and viscosity of the fluid (with respect to the surrounding gas, i.e., air); g is the gravity acceleration; and Vl is the terminal velocity to be determined. 664 M. Tembely et al. / Applied Thermal Engineering 31 (2011) 656e667 For a spherical fluid particle with a low Reynolds number, the Stokes flow analysis leads to the Hadamard-Rybczynski drag law, in which the shear stress on the surface induces an internal motion. The drag coefficient is expressed as follows: CD ¼ 8 2 þ 3k Rea 1 þ k (63) with the viscosity ratio k ¼ mf =ma . This result can be compared to that for a solid particle, where CD ¼ 24/Rea. Combining Eqs. (62) and (63), we deduce the terminal velocity Vl(V) (for a particle of volume V or diameter D), which is the relative velocity between the fluid particle and air: Vl ðVÞ ¼ 1 1 þ k D2 g  rf 6 2 þ 3k ma ra  (64) If we assume that the spray droplets reach velocities close to their terminal velocity before coalescing with arbitrary angles (b), the relative velocity between two particles can be expressed as follows: 2 Vl V 0 k ¼ Vl ðVÞ2 þVl V 0 u2r ¼ kVl ðVÞ 2 2Vl ðVÞVl V 0 cosb (65) Then, we can take the average velocity from the velocity directions RQ 0 to p/2, leading to hcosbi0;p=2 z2=p with hi0;Q ¼ ð1=QÞ 0 dx. 6.1.2. Collision frequency To assess the rate of collision or the collision frequency, we consider two particles of diameters D and D0 . When we consider the frame related to particle D, particle D0 evolves with a relative ! velocity of u r ðV; V 0 Þ. The necessary conditions to assure the collision between these particles are as follows: - Particle (D) is located in a cylinder with an axis that is parallel ! to u r ðV; V 0 Þ and with a diameter of D þ D0 , which also defines the cross-section Sc ¼ pðD=2 þ D0 =2Þ2 - To have a collision between times t and t þ dt, it is necessary that the distance between the centers of the two particles, measured ! parallel to u r ðV; V 0 Þ, is less than or equal to ur ðV; V 0 Þdt. In other words, the particle (D) must be located in the collision volume (i.e., the volume of cylindrical section Sc and length ur ðV; V 0 Þdt). Therefore, we deduce that the number of collisions of type D particles with a particle of type D0 during the time interval dt is np Sc ur ðV; V 0 Þdt. If we extend this result to all type D0 particles, the collision frequency can be expressed in the following form: Hf V; V 0 ¼ Sc np n0p VT ur ¼ pnp n0p VT V; V 0 2 D=2 þ D0 =2 ur V; V 0 (66) where np ; n0p are the numbers of particles per unit volume of size V; V 0 , respectively; and Sc is the collision cross-section of the two droplets. We approximate np n0p zðnðV; t ¼ 0ÞÞ=VT ÞðnðV 0 ; t ¼ 0ÞÞ=VT Þ using Eq. (55). With VT ¼ Vspray =aV , aV is a constant and expresses the spray volume fraction (or dispersed phase volume fraction), and Vspray is the total volume of the spray droplets (see the following subsection). Combining Eq. (57), (58) and (66), we obtain the coalescence kernel: 2 exp tcoal V; V =tcont V; V 0 6.2. Reformulation of the drop-size distribution equation The moments method and the size-binning method do not describe the precise behavior of the drop-size distribution. The Monte-Carlo method appears to be a more advantageous choice and describes the evolution of the drop-size distribution with precision. To carry out a precise numerical analysis, we reformulated the problem in terms of mass conservation and then developed a Mass Flow Algorithm (MFA). Multiplying Eq. (56) by the volume Vi and summing over all i leads to the mass conversation equation (assuming that the density of the fluid is constant): nc X Vi nði; tÞ ¼ i¼1 nc X nc X Mði; tÞ ¼ i¼1 Mði; 0Þ ¼ Vspray (68) i¼1 We can normalize this relationship by dividing with Vspray, the total volume of spray droplets, to obtain the following relation: nc X Mði; tÞ i¼1 Vspray ¼ nc X mði; tÞ ¼ 1 (69) i¼1 where we set mði; tÞ ¼ Mði; tÞ=Vspray . Due to mass conservation, we then have the following expression: nc d X mði; tÞ ¼ 0 dt i ¼ 1 (70) Multiplying by Vi and using the symmetry of kc (i, j), the following equation can be obtained: i 1 X vmði; tÞ ~c ði k ¼ vt j¼1 j; jÞmði mði; tÞ nc X j; tÞmðj; tÞ ~c ði; jÞmðj; tÞ k (71) jÞ ¼ Vspray kc ði ~c is jÞ=Vj . We can show that k j¼1 where we denote kc ði bounded, knowing KfDa expð Db Þ. Thus, we set KcN ¼ sup kc ði; jÞ. i;j1 In the continuous form, we can proceed by multiplying Eq. (51) R Vmax VNðtÞfn ðV; tÞdV for normalizaby V and dividing by Vspray ¼ Vmin tion purposes. By introducing the mass density function, MðV; tÞ ¼ VNðtÞf ðV; tÞ, we can obtain the Mass Flow Formulation, with gðV; tÞ ¼ MðV; tÞ=Vspray , using the symmetry of Kc: vg ðV; tÞ ¼ vt ZV ~c V K V 0; V 0 g V V 0 ; t g V 0 ; t dV 0 Vmin gðV; tÞ Vmax Z ~ c V; V 0 g V 0 ; t dV 0 K (72) Vmin ~ c ðV; V 0 Þ ¼ Vspray Kc ðV; V 0 Þ=V 0 and Kc is given by Eq. (67). where K The following condition is then assured by definition: ~ a CN DðVÞ=2 þ D0 V 0 =2 Vspray ur V; V 0 Kc V; V 0 zK 0 We can easily verify the following properties of the kernel: kc ði; jÞ ¼ kc ðj; iÞ  0. Eq. (67) can also be seen as a general form for a coalescence kernel that takes into account the mutual particle cross section, relative velocity and coalescence efficiency of particles. ð67Þ 2 where we set CN ¼ pðnðV; t ¼ 0ÞÞðnðV 0 ; t ¼ 0ÞÞ=Vspray Þ and the ~ a ¼ Ka aV . constant of our model K Vmax Z Vmin gðV; tÞdV ¼ 1 and d dt Vmax Z Vmin gðV; tÞdV ¼ 0 (73) M. Tembely et al. / Applied Thermal Engineering 31 (2011) 656e667 665 6.3. The Monte-Carlo scheme Once the problem is well-formulated, we can search for the drop-size distribution evolution using the Monte-Carlo scheme. A fixed time step Dt is chosen such that DtKcN < 1. We set tn ¼ nDt and mn(i) ¼ m(i, tn). The time is discretized using an explicit Euler scheme for i  1. Thus, we can compute mnþ1(i) using mass conservation (69) as follows: mnþ1 ðiÞ ¼ Dt i X ~ ði k c j; jÞmn ði jÞmn ðjÞ nc  X 1 j¼1 j¼1  Dt k~c ði; jÞ mn ðjÞmn ðiÞ (74) We can associate to fmn ðiÞ : 1  i  nc g the probability Pn defined on N*: nc X Pn ¼ (75) mn ðiÞdðiÞ i¼1 The following sequence for a set A3N* can be denoted by ðsA ðiÞÞi1 : sA ðiÞ ¼ 1 0 if i˛A otherwise (76) After some algebraic manipulations, we have the following relation: nc X mnþ1 ðiÞsA ðiÞ ¼ i¼1 nc X nc X fpði; jÞsA ði þ jÞ þ ð1 k¼1 j¼1 (77) pði; jÞÞsA ðiÞgmn ðiÞmn ðjÞ ~c ði; jÞ. The convergent Monte-Carlo Here we denote pði; jÞ ¼ Dt k scheme is then given by the following procedure. A quantity of N integers are chosen, and for all n  0 we approximate the solution at time tn by N particles that are denoted by iN;n ð1Þ; iN;n ð2Þ; .iN;n ðNÞ˛N* such that the following is true: ci N 1X sfig iN;n ðkÞ zmn ðiÞ N (78) k¼1 6.3.1. Initialization To initiate the computation, N numerical particles iN;0 ð1Þ; iN;0 ð2Þ; .iN;0 ðNÞ˛N* are chosen to yield the following expression: N 1X sfig iN;0 ðkÞ zm0 ðiÞ N Fig. 8. Comparison between analytical and numerical solutions of the second moment. extended to other configurations, such as for the breakup of spray droplets or evaporation. 6.4. Application to physically based drop-size distribution evolution The coupling with the previous physical approach from the Maximum Entropy Formalism (MEF) allows our model to be dynamic and sensitive to physical and operating conditions of the atomizer. Finally, we can deduce a physically based drop-size distribution evolution thanks to the given Monte Carlo scheme, using the following expression as the initialization step: m0 ðiÞ ¼ DVgðVi ; 0Þ ¼ iDVNFn ðiDDÞDD=Vspray (81) R Vmax where Vspray ¼ Vmin VFn dV. Finally, using Eq. (81) and with the Monte Carlo scheme given in Eq. (80), we can solve the evolution of the drop-size distribution for our initial distribution submitted to the coalescence effect using the MFA. 6.4.1. Monte-Carlo scheme convergence and validation Convergence of the model was tested using k(i, j) ¼ i þ j because an analytical solution is known with this kernel. As shown in Fig. 8, a convergence was obtained for a sample size of numerical particles (79) k¼1 6.3.2. Coalescence The sizes of particles at time tnþ1 are computed using the sizes of 1 ; X 2 ; .; X N particles at time tn. Let XN;n N;n N;n be N independent real random variables that are uniformly distributed in f1; 2.Ng and k ; 1  k  N be N independent real random variables that are UN;n uniformly distributed on [0,1]. Let us assume that all of the random k ; 1  k  N and U k ; 1  k  N are independent. The variables XN;n N;n new sizes of particles iN;nþ1 ðkÞ; 1  k  N are defined as follows: iN;nþ1 ðkÞ ¼ 8 < :      k k k iN;n ðkÞþiN;n XN;n if UN;n < p iN;n ðkÞ;iN;n XN;n iN;n ðkÞ otherwise (80) It can be shown that this scheme is convergent [2]. It should also be noted that the approach developed in this paper could easily be Fig. 9. Number based drop-size distribution, fn (D), and the coalescence effect. 666 M. Tembely et al. / Applied Thermal Engineering 31 (2011) 656e667 ~ a ¼ 0:5 and (b) K ~ a ¼ 5. Fig. 10. Number based drop-size distribution, fn (D), sensitivity to the coalescence effect (a) K of N ¼ 10000 and a number of time steps of P ¼ 400. The initial condition is as follows: f ði; 0Þ ¼ 1 0 if i ¼ 1 otherwise (82) The analytical expression is established in [32]; the second moment is given by the following relation: M2 ðtÞ ¼ N X i2 f ði; tÞ ¼ e2t (83) i¼1 6.4.2. Spray modeling Several tests were carried out to evaluate the evolution of the drop-size distribution. We observed that bigger drops appear in the spray over a period of time, as shown in Fig. 9. The first effect of coalescence was observed at a time of 5 ms. At longer times, we observed the coalescence effect with the emergence of bigger drops shifting toward the drop-size distribution in accordance with experimental measurements involving coalescence [33]. The dropsize distribution presented is bi-modal; such an observation has been made for ultrasonic spray [15,27], but so far no explanation has been advanced. In [27], during the analysis of the ultrasonic atomizer the presence of the second peak (near the main peak) was obtained. However, the second peak was neglected in the MEF distribution fitting. Although the existence of this supplementary peak is not fully confirmed for all ultrasonic atomization processes, based on the presented model, it is possible to retrieve and explain via coalescence this type of bi-modal distribution at the physical level. As shown in Fig. 10, the sensitivity study based on the coalescence effect leads to a physically based distribution (see Fig. 10(b)). This type of predictive capability, as well as the physically based distribution shape [33,34] of our model, is not achievable using a classical approach. It is worth noting that the temporal evolution of the model drop-size distribution could also be interpreted as the distribution at a spatial point zs such that zs ¼ Ust, where Us is the spray mean velocity and t is time. Our model could also be used for CFD modeling when the initial condition is a bi-modal (or even multi-modal) distribution, yielding a more physical basis unlike traditional approaches. An improvement of our scheme could be to adopt the Quasi-Monte-Carlo method (QMC) [35]. based formulation for spray modeling. Using the established Sauter Mean Diameter D32, Volume Mean Diameter D30, and SOD motion, we established a physically-based prediction model that couples the three-parameter generalized gamma distribution and conservation laws of MEF. The model parameters were deduced using a limited set of experimental measurements with the SOD. This new dynamic model is capable of predicting the drop-size distribution for well-specified operating conditions, fluid and nozzle structure properties. This new approach avoids the traditional adjustment for each operating condition and has better predictive capabilities. The population balance equation was analyzed, taking into account the interactions between droplets. Thus, we established the evolution and resolution of the drop-size distribution equation submitted to the coalescence effect. In contrast to Direct Numerical Simulation (DNS), the Mass Flow Algorithm (MFA) chosen here enables preservation of the total mass of particles. In this work, we considered only binary interactions where two droplets can coalesce to form a bigger one prior to impact with the substrate. Based on a physical hypothesis, we used our proposed coalescence kernel and coupled the model with our previous physically based approach. To solve the problem, a Monte-Carlo Method, which was shown to be convergent, was developed to highlight the formation of new drops due to coalescence, leading to a physically based bi-modal distribution. As a future perspective, it is possible to improve this method by adopting a quasi-Monte-Carlo simulation method. This would consist of replacing the (random) Monte-Carlo simulation algorithm by a deterministic one. Nomenclature a b fp g N Ni rf mf ab sf qE semi-major axis length of the nozzle exit shape semi-minor axis length of the nozzle exit shape vibration frequency of the piezoceramic disk gravity acceleration total number of drops number of drops in each class i fluid density fluid viscosity nozzle tip beveled angle fluid surface tension equilibrium contact angle fluid/structure 7. Conclusion References In this paper, we have performed a theoretical study of the instantaneous drop-size distribution and its temporal evolution applied to a new SOD device. From two approaches of the MEF (i.e., specific and general formulations), we derived a new physically [1] P. Calvert, Inkjet printing for materials and devices, Chem. Mater. 13 (2001) 3299. [2] M. Tembely, Atomization Induced by Fluid-Structure Interactions, PhD thesis, Université de Grenoble, 2010. M. Tembely et al. / Applied Thermal Engineering 31 (2011) 656e667 [3] A.H. Lefebvre, Atomization and Sprays. Hemisphere Publishing Corporation, New York, NY, USA, 1989. [4] G.G. Nasr, A.J. Yule, L. Bendig, Industrial Sprays and Atomization: Design, Analysis and Applications. Springer, 2002. [5] J. Eggers, E. Villermaux, Physics of liquid jets, Rep. Prog. Phys. 71 (03660) (2008). [6] A.L. Yarin, Free Liquid Jets and Films: Hydrodynamics and Rheology. Longman Scientific & Technical and Wiley & Sons, Harlow, New York, 1993. [7] E. Villermaux, Fragmentation, Annu. Rev. Fluid Mechanics 39 (2007) 419e446. [8] E. Villermaux, P. Marmottant, J. Duplat, Ligament-mediated spray formation, Phys. Rev. Lett. 92 (074501) (2004). [9] E. Babinski, P.E. Sojka, Modeling drop size distributions, Prog. Energy Combust. Sci. 28 (2002) 303e329. [10] M. Faraday, On the forms and states assumed by fluids in contact with vibrating elastic surfaces, Phil. Trans. R. Soc. Lond. 52 (1831) 319e340. [11] R.J. Lang, Ultrasonic atomisation of liquids, J. Acoust. Soc. Am. 34 (1962) 6e8. [12] F. Lacas, P. Versaevel, P. Scouflaire, G. Coeur-Joly, Design and performance of an ultrasonic atomization system for experimental combustion applications, Part. Part. Syst. Charact. 11 (1994) 166e171. [13] D.M. Kirpalani, F. Toll, Revealing the physicochemical mechanism for ultrasonic separation of alcohol-water mixtures, J. Chem. Phys. 117 (2002) 3874e3877. [14] P. Chen, Nonlinear wave dynamics in Faraday instabilities, Phys. Rev. E 65 (036308) (2002). [15] M. Dobre, Caractérisation stochastique des sprays ultrasoniques: le formalisme de l’entropie maximale, Thesis, UCL, 2003. [16] H. Liu, Science and Engineering of Droplets - Fundamentals and Applications. William Andrew Publishing-Noyes, 2000. [17] X. Li, R.S. Tankin, Derivation of droplet size distribution in sprays by using information theory, Comb. Sci. Technol. 60 (1988) 345e357. [18] C. Dumouchel, The maximum entropy formalism and the prediction of liquid spray drop-size distribution, Entropy 11 (2009) 713e747. [19] J. Cousin, S.J. Yoon, C. Dumouchel, Coupling of classical linear theory and maximum entropy formalism for prediction of drop-size distribution in sprays: application to pressure swirl atomizers, Atomization and Sprays 6 (1996) 601e622. [20] J.H. Lienhard, P.L. Meyer, A physical basis for the generalized gamma distribution, Quart. Appl. Math. 25 (1967) 330e334. 667 [21] C. Dumouchel, A new formulation of the maximum entropy formalism to model liquid spray drop-size distribution, Part. Part. Syst. Charact. 23 (2006) 468e479. [22] N. Agmon, Y. Alhassid, R.D. Levine, An algorithm for finding the distribution of maximal entropy, J. Comp. Phys. 30 (2) (1977) 250e258. [23] M. Lecompte, C. Dumouchel, On the capability of the generalized gamma function to represent spray drop-size distribution, Part. Part. Syst. Charact. 25 (2008) 154e167. [24] J.N. Kapur, Twenty-five years of maximum-entropy methods, J. Math. Phy. Sci. 17 (1983) 103e156. [25] N. Fdida, Développement d’un système de granulométrie par imagerie. Application aux sprays larges et hétèrogènes, PhD thesis, Coria, 2008. [26] J.B. Blaisot, J. Yon, Droplet size and morphology characterization for dense sprays by image processing: application to the diesel spray, Exp. Fluids 39 (2005) 977e994. [27] C. Dumouchel, D. Sindayihebura, L. Bolle, Application of the maximum entropy formalism on sprays produced by ultrasonic atomizers, Part. Part. Syst. Charact. 20 (2003) 150e161. [28] S. Chapman, T.G. Cowling, The Mathematical Theory of Non-Uniform Gases. Cambridge University Press, Cambridge, 1970. [29] G. Kocamustafaogullari, M. Ishii, Foundation of the interfacial area transport equation and its closure relations, Int. J. Heat Mass 38 (1995) 481e493. [30] A.K. Chesters, G. Hoffman, Bubble coalescence in pure liquids, Appl. Sci. Res. 38 (1982) 353e361. [31] V.G. Levich, Physicochemical Hydrodynamics. Prentice-Hall, Englewood Cliffs, 1962. [32] A. Golovin, The solution of the coagulating equation for cloud droplets in a rising air current, Izv. Geophys. Ser. 5 (1963) 482e487. [33] S.S. Yoon, J.C. Hewson, P.E. Desjardin, D.J. Glaze, A.R. Black, R.R. Skaggs, Numerical modeling and experimental measurements of a high speed solidcone water spray for use in fire suppression applications, Int. J. Multiphase Flow 30 (11) (2004) 1369e1388. [34] G. Brenn, S.t. Kalenderski, I. Ivanov, Investigation of the stochastic collisions of drops produced by Rayleigh breakup of two laminar liquid jets, Phys. Fluids 9 (1997) 349. [35] C. Lécot, W. Wagner, A quasi-Monte Carlo scheme for Smoluchowski’s coagulation equation, Math. Comput. 73 (2004).