+ All Categories
Home > Documents > A Study on Heat Transfer in Dropwise Condensation

A Study on Heat Transfer in Dropwise Condensation

Date post: 08-Jan-2022
Category:
Upload: others
View: 5 times
Download: 0 times
Share this document with a friend
158
The Pennsylvania State University The Graduate School A Study on Heat Transfer in Dropwise Condensation A Dissertation in Mechanical Engineering by Sanjay Adhikari © 2020 Sanjay Adhikari Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy August 2020
Transcript
Page 1: A Study on Heat Transfer in Dropwise Condensation

The Pennsylvania State University

The Graduate School

A Study on Heat Transfer in Dropwise

Condensation

A Dissertation in

Mechanical Engineering

by

Sanjay Adhikari

© 2020 Sanjay Adhikari

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

August 2020

Page 2: A Study on Heat Transfer in Dropwise Condensation

ii

The dissertation of Sanjay Adhikari was reviewed and approved by the following:

Alexander S. Rattner

Assistant Professor of Mechanical Engineering

Dissertation Adviser

Chair of Committee

Fan-Bill Cheung

George L. Guillet Professor of Mechanical and Nuclear Engineering

Tak-Sing Wong

Wormely Family Early Career Professor in Engineering

Assistant Professor of Mechanical Engineering

Kristen Fichthorn

Merrell Fenske Professor of Chemical Engineering and Professor of Physics

Daniel Haworth

Associate Head for Graduate Programs, Interim

Page 3: A Study on Heat Transfer in Dropwise Condensation

iii

Abstract

Dropwise condensation (DWC) is a high intensity heat transfer phenomenon, with heat fluxes up to an order

of magnitude higher than in filmwise condensation. Many industrial applications like desalination, thermal

management, power generation and fog water harvesting may benefit from employing DWC. Considerable work

has been done in this field and yet there are some challenges which remain open for research and further

investigation. High resolution numerical studies that capture the complex hydrodynamics of larger droplets as they

coalesce and move on the surface, effectively refreshing it for renucleation, have still not been realized as the drop

size distribution spans six orders of magnitude. A new Volume of Fluid (VOF) based multi-scale approach, which

resolves the grid scale larger droplets O(50 μm) and models the heat transfer for the subgrid scale smaller droplets,

is proposed in this dissertation. To supplement this approach, heat transfer models for individual droplets (Chapter

1) and droplet interaction (Chapter 2) are presented. Modeling work described in chapter 2 also solves many

challenges associated with resolving grid scale droplets in VOF simulations. The proposed multi-scale approach

also requires a transient heat transfer closure model for the subgrid scale droplets. Experimental studies have so far

focused only on the steady state behavior of DWC. The startup period is still not well understood. Transient

phenomena in DWC startup may significantly affect the performance of proposed applications like vapor chambers

subjected to pulsed heat fluxes, or refrigeration cycles operating intermittently. This dissertation provides first

characterization of fast transients during startup (chapter 4). A unique experimental facility and instrumentation are

developed for measurements. Inverse numerical methods are employed to extract data (q" and HTC) from these

measurements. Three distinct phases during DWC startup have been identified, which further highlight the

importance of characterizing this period. The experimental data can be coupled with the numerical work described

in this dissertation to formulate a robust, widely applicable simulation framework.

Page 4: A Study on Heat Transfer in Dropwise Condensation

iv

This dissertation is based upon the work partially funded by the United States National Science Foundation under

the Award Number CBET-1652578. This report was prepared as an account of work sponsored by an agency of the

United States Government. Neither the United States Government nor any agency thereof, nor any of their

employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy,

completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use

would not infringe privately owned rights. References herein to any specific commercial product, process, or

service-water by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its

endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and

opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or

any agency thereof.

Page 5: A Study on Heat Transfer in Dropwise Condensation

v

Table of Contents

List of Figures ........................................................................................................................................................ viii

List of Tables .......................................................................................................................................................... xiii

Acknowledgements ................................................................................................................................................ xiv

Chapter 1 Introduction and Literature Review ......................................................................................................1

1.1 Dropwise Condensation: Application and Mechanisms .............................................................................2

1.2 Prior Research in Dropwise Condensation .................................................................................................7

1.2.1 Major Advances in the History of Dropwise Condensation ...............................................................7

1.2.2 Heat Transfer Models of Dropwise Condensation .............................................................................7

1.2.3 Computational Approaches to Modeling Dropwise Condensation ..................................................10

1.2.4 Goals of the Present research and Dissertation .................................................................................13

Chapter 2 Detailed Characterization of Individual Droplet Heat Transfer .........................................................17

2.1 Modeling Approach ..................................................................................................................................19

2.2 Simulation Studies ....................................................................................................................................21

2.3 Results and Discussion .............................................................................................................................24

2.3.1 Proposed Correlation ........................................................................................................................24

2.3.2 Validation of Results ........................................................................................................................25

2.3.3 Internal temperature Distribution in Droplets ...................................................................................29

2.3.4 Implications of varying Conduction Resistance Factor in Dropwise Condensation ........................31

2.4 Conclusion ................................................................................................................................................33

Chapter 3 Characterizing Droplet Interactions During Coalescence ..................................................................35

3.1 Modeling Approach ..................................................................................................................................39

3.1.1 Governing equations .........................................................................................................................39

3.1.2 Surface Tension Model .....................................................................................................................42

3.1.3 Phase Change Model ........................................................................................................................43

3.1.4 Dynamic Contact Angle Model ........................................................................................................44

3.1.5 Solution Algorithm ...........................................................................................................................45

3.2 Simulation Studies ....................................................................................................................................47

3.2.1 Studied Domain and Case setup .......................................................................................................47

Page 6: A Study on Heat Transfer in Dropwise Condensation

vi

3.2.2 Grid Resolution, Sensitivity Study and Dynamic Contact Angle Validation ...................................51

3.3 Validation of Results ................................................................................................................................54

3.3.1 Results and Discussion .....................................................................................................................57

3.3.2 Characterizing Coalescence ..............................................................................................................59

3.3.3 Effect of Coalescence on Heat Transfer ...........................................................................................64

3.3.4 Simulation Data and Proposed Correlation for Droplet Coalescence Time Constant ......................69

3.4 Conclusions ..............................................................................................................................................71

Chapter 4 Heat Transfer Measurements of Fast Transient Startup of Dropwise Condensation..........................73

4.1 Experimental Design and Setup ...............................................................................................................76

4.1.1 Apparatus ..........................................................................................................................................77

4.1.2 Test Section ......................................................................................................................................80

4.1.3 Sensor Plate and Amplification Module ...........................................................................................84

4.1.4 Dielectric Thermal Conductivity Measurement ...............................................................................88

4.2 Results and Discussion .............................................................................................................................92

4.2.1 Transient Heat Transfer Measurements ............................................................................................92

4.2.2 Heat Flux and Heat Transfer Coefficient Calculation Methodology ................................................94

4.2.3 Validation Using Constriction Resistance and Theoretical Drop-Size Distribution Model .............98

4.2.4 Discussion on Transient Heat Fluxes and Heat Transfer Coefficients ...........................................104

4.2.5 Conclusions ....................................................................................................................................110

Chapter 5 Preliminary Study on Multi-Scale Simulation of Dropwise Condensation ......................................111

5.1 Modeling Approach ................................................................................................................................114

5.1.1 Grid-Scale Modeling ......................................................................................................................114

5.1.2 Subgrid Scale Modeling .................................................................................................................115

5.2 Illustrative Domain and Case Setup .......................................................................................................117

5.3 Preliminary Results and Discussion .......................................................................................................119

5.4 Conclusions ............................................................................................................................................122

Chapter 6 Conclusions and Recommendations for Future Research ................................................................123

6.1 Numerical Simulations of dropwise condensation .................................................................................124

6.2 Experimental Measurement of Heat Transfer During Dropwise Condensation Startup ........................125

6.3 Future Research Work ............................................................................................................................126

Page 7: A Study on Heat Transfer in Dropwise Condensation

vii

Appendix ................................................................................................................................................................128

A 1 Calibrating the Test Plate RTDs .........................................................................................................128

A 2 Population Balance Theory ................................................................................................................130

Bibliography ...........................................................................................................................................................134

Page 8: A Study on Heat Transfer in Dropwise Condensation

viii

List of Figures

Figure 1.1 a) Schematic of the Droplet with radius r, contact angle θ, conduction resistance Rcond and interfacial

resistance Rint. b) Magnified view of the interfacial resistance ..................................................................................4

Figure 1.2 Comparison between droplet covered surfaces a) From the numerical study of Burnside and Hadi [1];

b) Proposed multi-scale simulation test surface. ..................................................................................................... 13

Figure 2.1 Representative mesh sizes: (a) coarse Δ3 = 2.2 × 102mm; (b) fine Δ2 = 1.2 × 10 − 2mm; (c) finest

Δ1 = 6.0 × 10 − 3mm ........................................................................................................................................... 21

Figure 2.2 𝑓𝐵𝑖, 𝜃 vs. Bi-1 for different contact angles. ............................................................................................ 23

Figure 2.3 Comparison of absolute values of f predicted with proposed correlation and conduction simulations. 25

Figure 2.4 Nu vs. Bi for different contact angles (50° to 80°). ............................................................................... 26

Figure 2.5 Non-dimensional temperature 𝑻−𝑻𝒘𝒂𝒍𝒍

𝑻𝒔𝒂𝒕−𝑻𝒘𝒂𝒍𝒍 distribution inside a droplet with varying Bi. a) Bi = 1000, b)

Bi = 2 and c) Bi = 0.2. ............................................................................................................................................. 29

Figure 2.6 Heat flux along the condensing surface of the droplet with 𝜽 = 𝟗𝟎° for different Bi. 𝜷 is the angle

subtended by any point on the droplet surface and the horizontal (inset). .............................................................. 30

Figure 2.7 Non-dimensional temperature distribution inside droplets with different contact angles for Bi = 1000.

a) 𝛉 = 𝟒𝟎°, b) 𝛉 = 𝟗𝟎° and c) 𝛉 = 𝟏𝟒𝟎°. .............................................................................................................. 31

Figure 2.8 Comparison of droplet growth between OpenFoam VOF simulation, theoretical relation with f obtained

using the correlation and theoretical relation with f = 0.25. .................................................................................... 33

Figure 3.1 Coalescing droplet mid-plane temperature profiles. Both parent droplet radii are 𝟐𝟓 𝛍𝐦. The

maximum velocity observed during coalescence is 𝐔𝐦𝐚𝐱 ~ 𝟑 𝐦 𝐬 − 𝟏. .............................................................. 38

Page 9: A Study on Heat Transfer in Dropwise Condensation

ix

Figure 3.2 Advancing and receding contact angles measured using volume changing technique (liquid injection or

removal from the top of the droplet). a) Increasing volume of the drop. b) Decreasing volume of the drop. ......... 45

Figure 3.3 Phase change volume-of-fluid solver algorithm flow chart. .................................................................. 46

Figure 3.4 Representative domain for the VOF coalescence study. The face showing the droplets is the symmetry

plane; the bottom surface, which is not visible in the figure, is the condensing surface; and the remaining 4 faces

are open atmosphere. ............................................................................................................................................... 47

Figure 3.5 Steady state temperature distribution on the droplet surfaces before merging. ..................................... 49

Figure 3.6 Representative droplet isosurfaces, indicating VOF mesh resolutions for the droplets from the

parametric study. a) For smaller droplet radius 𝒓𝟏 = 𝟏 𝛍𝐦 and radius ratio Rt = 1. b) For 𝒓𝟏 = 𝟏 𝛍𝐦 and Rt = 2.

c) For 𝒓𝟏 = 𝟏 𝛍𝐦 and Rt = 4.................................................................................................................................. 51

Figure 3.7 Heat transfer grid independence study (Q vs 𝒏𝒓 where 𝒏𝒓 = # 𝒄𝒆𝒍𝒍𝒔/𝒓𝟏) for 𝒓𝟏 = 𝟏 𝛍𝐦 and 𝒓𝟏 =

𝟐𝟓 𝛍𝐦. .................................................................................................................................................................... 52

Figure 3.8 Colliding droplet contact patch diameter on surface (D) vs. time (t) for different grid resolutions and

experimental data [121]. .......................................................................................................................................... 53

Figure 3.9 Comparison of the simulation data (heat transfer rate (Q) vs. Biot number (Bi)) with that of Sadhal and

Martin [91], and conduction-only finite element COMSOL simulations. ............................................................... 55

Figure 3.10 Representative mesh sizes: (a) coarse 𝚫𝟑 = 𝟏. 𝟗𝟕 × 𝟏𝟎 − 𝟗𝛍𝐦− 𝟏. 𝟗𝟗 × 𝟏𝟎 − 𝟓𝛍𝐦; (b) fine 𝚫𝟐 =

𝟓. 𝟖𝟒 × 𝟏𝟎 − 𝟏𝟎𝛍𝐦− 𝟓.𝟗𝟎 × 𝟏𝟎 − 𝟔𝛍𝐦; (c) finest 𝚫𝟏 = 𝟏. 𝟒𝟖 × 𝟏𝟎 − 𝟏𝟎 𝛍𝐦− 𝟏. 𝟓𝟎 × 𝟏𝟎 − 𝟔𝛍𝐦. ........ 56

Figure 3.11 Total heat transfer rate (Q) vs. droplet radius (r) for water condensation at atmospheric pressure. The

dashed lines indicate the rate of variation of heat transfer with droplet radius for different size droplets. ............. 60

Figure 3.12 . Comparison of a) condensing droplet heat transfer rate vs. time, b) base area vs. time, and c) contact

line (perimeter) length vs. time for two merging drops of radii 𝟔. 𝟐𝟓 𝛍𝐦 (𝐁𝐢 ≈ 𝟎. 𝟑)........................................... 62

Page 10: A Study on Heat Transfer in Dropwise Condensation

x

Figure 3.13 Decaying exponential function superimposed over Q vs Time plot of two equal merging drops of radii

𝟔. 𝟐𝟓 𝛍𝐦 each. ........................................................................................................................................................ 63

Figure 3.14 Differences between subgrid scale heat transfer due to microscale condensing droplets around two

coalescing r = 25 μm droplets: a) the dynamic coalescence case; and b) the static coalescence case. Top-down

view snapshots of the droplet shapes, represented in black, are presented. The colored area underneath the

droplets shows the evolution of subgrid scale heat flux due to small condensing droplets around the primary

coalescing droplets 𝐪"𝐒𝐆𝐒 ∗ 𝐭 = (𝐪"𝐒𝐆𝐒 −𝐦𝐢𝐧(𝐪"𝐒𝐆𝐒(𝐭)))/(𝐦𝐚𝐱(𝐪"𝐒𝐆𝐒(𝐭)) −𝐦𝐢𝐧(𝐪"𝐒𝐆𝐒(𝐭))). ............... 67

Figure 3.15 Cumulative heat transfer for an interface tracked dynamic coalescence case and a static case. Initially,

for both cases, both droplets have radius 𝒓 = 𝟐𝟓 𝛍𝐦. ............................................................................................ 68

Figure 3.16 Dimensionless time constant (𝝉 ∗) vs. dimensionless smaller droplet radius (𝒓𝟏 ∗) and droplet radius

ratio Rt. The curves are generated using the correlation given in Eqn. 3.22 and the simulation data is represented

by the markers. ........................................................................................................................................................ 70

Figure 4.1 Experimental setup to capture transient heat transfer data during DWC. a) Schematic of the setup, b)

Photograph of the full test stand, and c) photographs of major components: (1) Steam generator; (2) Test section;

(3) Test plate with resistance temperature detectors (RTDs) printed on both sides and heater for drying between

tests; (4) Full cone spray nozzle; (5) Primary pressurized cooling water tank; (6) Auxiliary water collection tank;

(7) Venting and vacuum arrangement for steam; (8) Peripheral heater for the test plate; (9) constant current source

for RTDs and signal amplifier; (10) Data acquisition system (DAQ); (11) View-port for camera. ........................ 78

Figure 4.2 Test Section: 3 – Test plate; 4 – Full cone spray nozzle; 2.1 – Coolant supply line; 2.2 – Coolant

discharge port to the auxiliary tank; 2.3 – Condensate return port; 2.4 – Steam inlet; 2.5 – Sight glass assembly;

2.6 – Port for venting, vacuum, and pressure transducer; 2.7 – Temperature sensor port. (Dimensions in mm) ... 80

Figure 4.3 a) Evolution of steam temperature 𝐓𝐄 with time; b) Evolution of steam pressure 𝐏𝐄 with time. ......... 81

Page 11: A Study on Heat Transfer in Dropwise Condensation

xi

Figure 4.4 a) High speed camera arrangement; b) Camera view during the experimental run. .............................. 83

Figure 4.5 Test plate: 1 – Gold-plated copper track used as an RTD in a 4-wire configuration; 2 – Gold finger

contacts for connection to the RTD processing circuits; 3 – Area visible to the camera (Figure 4.4-b). ................ 84

Figure 4.6 a) Test plate dimensions in 𝐦𝐦. b) Stack-up of layers over the aluminum substrate on the condensing

side of the plate: (1) 100 nm thick conformal coat of parylene-C; (2) 80 nm thick (nominal) ENIG® finish layer

(3) 8 𝛍𝐦 thick copper tracks; (4) 85 𝛍𝐦 thick COBRITHERM (Benmayor Aismalibar. proprietary) dielectric

layer. ........................................................................................................................................................................ 85

Figure 4.7 a) Constant current supply and signal amplification board: 1 – Edge connector, which mates with the

test plate terminals; 2 – Signal amplification block; 3 – Constant current supply block; 4 – Signal cable connector

to the DAQ; b) Schematic of a single RTD, current and amplifier loop: 1 – Constant current supply; 2 – Voltage

signal amplification. ................................................................................................................................................ 87

Figure 4.8 Conduction resistance network schematic between surface temperature measurements on the cold and

the condensing faces of the test plate. ..................................................................................................................... 89

Figure 4.9 COMSOL® Multiphysics mesh to determine thermal conductivity of the dielectric layer: 1 – Coarse

mesh; 2 – Fine mesh; 3 – Finest mesh. .................................................................................................................... 90

Figure 4.10 DWC on the vertical test plate at different times after startup (pressure = 1 atm) ............................... 93

Figure 4.11 Schematic representing 1 dimensional discretization of the test plate, which is divided into separate

material zones with different material properties and discretization lengths: 1 – Dielectric zone (discretization

length 𝚫𝐱𝐝); 2 – Aluminum zone (discretization length 𝚫𝐱𝐀𝐥); 3 – Dielectric zone (discretization length 𝚫𝐱𝐝); 4 –

Boundary; 5 – Aluminum-dielectric interface; 6 – Aluminum-dielectric Interface; 7 – Boundary. ....................... 94

Figure 4.12 Temperature profile through the plate for time t = 0.3 s. ..................................................................... 97

Page 12: A Study on Heat Transfer in Dropwise Condensation

xii

Figure 4.13 Schematic showing the constriction resistance 𝐑𝐬 and the promoter layer resistance 𝐑𝐂 between the

surface temperature measured by the RTDs on the condensation side (𝐓𝐬𝐮𝐫𝐟) and the Steam temperature (𝐓𝐄). 99

Figure 4.14 a) Surface temperature on the Condensing side 𝐓𝐬𝐮𝐫𝐟, surface temperature on the cooling side 𝐓𝐂

and the steam temperature 𝐓𝐄 during the experiment: 1 – First inflection point; 2 – Second inflection point; b)

Magnified presentation of the chart area indicated above it. ................................................................................. 105

Figure 4.15 Solved heat fluxes on the cooled and condensing faces of the test plate. .......................................... 106

Figure 4.16 Evolution of heat transfer coefficients on the condensing face (a, 𝐇𝐓𝐂𝐃𝐖𝐂) and cooled face (b,

𝐇𝐓𝐂𝐂). ................................................................................................................................................................... 107

Figure 4.17 Schematic cross-section representing the test plate RTD calibration setup: (1) Reference temperature

probe; (2) High thermal conductivity aluminum block; (3) intermediate thermal conductivity stone block; (4)

Temperature controlled heated plate; (5) – Thermal insulation; (6) Signals to data acquisition system; (7) Test

plate. ...................................................................................................................................................................... 128

Figure 5.1 Illustrative distribution of grid scale and subgrid scale drop sizes. Grid scale droplets are resolved using

VOF, and an average heat flux model is applied instead of resolving subgrid scale droplets............................... 113

Figure 5.2 Average heat transfer coefficient vs time of a surface (given by Glicksman and Hunt [28]). ............. 116

Figure 5.3 Representative VOF domain for multi-scale simulation test case. ...................................................... 117

Figure 5.4 a) Liquid coverage on the condensing surface after the elapse of 1 minute in the multi-scale simulation

b) 3D rendering of another test simulation with contact angle set to 𝟏𝟎𝟎°. ......................................................... 119

Figure 5.5 Cumulative surface coverage by the grid scale droplets, of the condensing surface vs droplet radii for

different times. ....................................................................................................................................................... 120

Figure 5.6 a) Grid scale heat transfer map; b) Subgrid scale heat transfer map. ................................................... 121

Page 13: A Study on Heat Transfer in Dropwise Condensation

xiii

List of Tables

Table 2.1 Coefficients for the correlation given in Eqn. 2.10. ................................................................................ 24

Table 2.2 Average heat flux through the base of droplets with representative hydrophobic Bi and 𝜃 values,

computed with VOF solver, COMSOL conduction studies, and proposed correlation. ......................................... 28

Table 3.1 Boundary conditions for the VOF simulation domain ............................................................................ 48

Table 3.2 Simulation test matrix for the parametric study of coalescing droplets, with varying smaller radius (r1)

and the ratio between the radii of the merging droplets (Rt). Time constant 𝜏 is obtained from the simulation

results. ...................................................................................................................................................................... 58

Table 3.3 Constants for the droplet coalescence time constant correlation (Eqn. 3.22) .......................................... 69

Table 3.4 𝝉 for representative test cases obtained from best-fit exponential curves to the experimental data and the

proposed correlation (Eqn. 3.22). ............................................................................................................................ 70

Table 4.1 Grid independence calculation summary for the 1D, unsteady conduction equation solver. .................. 97

Table 4.2 Comparison between theoretically calculated heat flux and heat transfer coefficient with mean steady

state quantities from the experiment. ..................................................................................................................... 103

Table 5.1 Boundary conditions for the VOF domain. ........................................................................................... 118

Page 14: A Study on Heat Transfer in Dropwise Condensation

xiv

Acknowledgements

First of all, I would like to thank my advisor Dr. Alexander S. Rattner who has guided me through the

entire period of my PhD. He has been patient and extremely kind. I will forever be grateful and indebted to

him. I would also like to thank all my past and present lab mates, especially Mahdi Nabil and Chris Greer, for

helping me out whenever I needed it.

I appreciate the financial support of this project by the U.S. National Science Foundation and

Department of Mechanical Engineering at The Pennsylvania State University during the past five years.

I am really thankful to my committee members: Dr. Fan-Bill Cheung, Dr. Tak-Sing Wong and Dr.

Kristen Fichthorn.

Finally, I would like to thank my family, Anand, Ashley, Hema and Sachin, for their unwavering love

and support.

Page 15: A Study on Heat Transfer in Dropwise Condensation

1

Chapter 1

Introduction and Literature Review

Page 16: A Study on Heat Transfer in Dropwise Condensation

2

1.1 Dropwise Condensation: Application and Mechanisms

Dropwise condensation has been identified as a promising heat transfer mechanism that can sustain

heat fluxes up to an order of magnitude higher than filmwise condensation [1]–[5]. It could play an important

role in a number of practical applications, including power generation condensers [6], desalination [7], and

thermal management [8]. For example, Brunt and Minken [9], in their 1958 study on seawater evaporative

desalination, found a 50% increase in the output of such a system using dropwise condensation. Recently,

investigations have also explored ambient fresh water harvesting using dropwise condensation (e.g., fog water

collection [10], [11] and water collection in arid/desert conditions [12]).

The dropwise condensation process is initiated from minute (<1 μm diameter) primary droplets at

discrete nucleation sites (often 107 − 109 sites mm−2), which are distributed over the condensing surface [13].

If the radius of a primary droplet is greater than the smallest thermodynamically stable value [14], it will grow

due to condensation and coalescence with neighboring droplets until it is large enough to be removed by body

forces (e.g., gravity). Larger droplets slide during coalescence and departure, clearing portions of the cooled

surface, and allowing the formation of new primary droplets at the nucleation sites.

After the initiation of dropwise condensation, the surface state evolves towards a quasi-steady condition

[2], [14]. A final drop size distribution is eventually obtained when the mean rate of condensation is balanced

by the rate of liquid removal due to departing droplets. Moreover, the number distribution (Area fraction

covered by droplets of each size range) is bimodal [15] with large droplets 𝑂(10−3) m and very small droplets

𝑂(10−6) m. The problem of predicting the steady state heat transfer rate in dropwise condensation can therefore

be divided into two parts: (1) predicting the heat transfer through individual droplets of varying sizes, and (2)

quantifying the distribution of drop sizes over the condenser surface. The effects of sliding, departing and

Page 17: A Study on Heat Transfer in Dropwise Condensation

3

merging droplets is a localized phenomenon on a steady state surface, and a time and area averaged heat transfer

coefficient should also capture such effects. In an individual droplet, the heat of condensation transfers through

the liquid-vapor interface (R"int) and then conducts through the base of the droplet (Rcond) to the cooled surface.

Figure 1.1 illustrates the heat transfer process in a droplet. Curvature of the droplet at the liquid-vapor interface

results in lowering of the saturation temperature at the interface [16]–[18]. Many studies have claimed that the

effect of curvature and interfacial resistance is negligible [19]–[21] as compared to conduction resistance in the

droplet. However, it is shown in Chapter 2 that interfacial resistance effect is not negligible in typical

condensation conditions. Eqn. 1.1 [13] present one of the most widely used analytic models for the average

heat flux q” through a steady state surface and takes into account Rcond, R"int, curvature effect and droplet

distribution.

q" =

1

3r23

×∫

{

ΔT −2σTsathlvρr

K1rk+ K2 (

0.6270.664)

Tsathlv2 ρv

(γ + 1)(γ − 1)

[RTsat2π ]

2

}

r−23dr

r

r

1.1

Here, r is the maximum droplet radius (𝑂)~10−3m [22] for water vapor condensing at Patm, r is the

minimum thermodynamically viable radius (𝑂)~10−9m [22], ΔT is the temperature difference between the

surface and the vapor, σ is the surface tension, Tsat is the saturation temperature, hlv is the enthalpy of

vaporization, ρ is the liquid density, K1 and K2 are constants. K1 and K2 are typically assumed as constants close

to unity [13]. ρv is the vapor density, γ is the ratio of the principal specific heat capacities (cp,v/cv,v), and R is

the specific ideal-gas constant. The second term in the numerator of the integrand is due to curvature of the

droplet. The first term in the denominator of the droplet is due to conduction resistance, and the second term is

due to interfacial resistance. The drop size distribution generally used in Eqn. 1.1 is given in Eqn. 1.2 [14].

Page 18: A Study on Heat Transfer in Dropwise Condensation

4

Figure 1.1 a) Schematic of the Droplet with radius r, contact angle θ, conduction resistance

Rcond and interfacial resistance Rint. b) Magnified view of the interfacial resistance

The drop size distribution on a surface is a critical component of heat transfer models. It has been shown

that heat transfer through a droplet of radius r can be integrated over the drop size distribution to predict overall

heat flux through a surface in steady state [22]–[24]. Several experimental [23], [25], [26], numerical [1], [27]–

[30], and theoretical [2], [14], [17] studies have helped in developing a theory in drop size distribution over a

surface during steady state dropwise condensation. Some key points are described as follows.

Eqn. 1.2 [17] gives the fraction (χ) of the surface covered by the drops having radii in the range of r to

r.

Page 19: A Study on Heat Transfer in Dropwise Condensation

5

χ = χ (r

r)

χ(1) = 0

1.2

The above form is supported by the observation that photograph of a surface undergoing dropwise

condensation appear the same under different magnifications (self-similar) [22], [30]–[34]. Le Fevre and Rose

[14] have proposed the following relation for 𝜒, which is supported by data from the experimental studies [23],

[25], [26].

χ = 1 − (r

r)

13

1.3

Heat transfer through a droplet q"(r) of radius r can be combined with the drop size distribution to

predict average heat flux through a surface [17].

q" = ∫ q"(r) × (−χ′ (

r

r))dr

r

r

r

1.4

1.1.1 Transient Heat Transfer Modeling

q"(r) has recently been completely characterized by Chavan et al. [35] (hydrophobic surfaces) and

Adhikari et al. [36] (hydrophilic and hydrophobic surfaces). The latter study in conjunction with an evolving

transient drop size distribution, and using the approach given by Eqn. 1.4, can be utilized to predict the overall

transient heat flux through a surface. However, Limited transient drop size distribution [28] or heat flux [28],

[37] data are available in literature.

Page 20: A Study on Heat Transfer in Dropwise Condensation

6

Most of the experiments in literature have employed temperature sensors in the substrate, under the

condensing surface, at varying depths[14], [26], [38]–[40]. Surface temperature can be extrapolated from such

measurements using Fourier’s law. However, this approach may be difficult to adapt to transient heat transfer

measurements of DWC startup. During startup, the temperature profile through the substrate will be non-linear,

making instantaneous heat flux estimation difficult. Further, conventional DWC heat transfer test sections have

relatively thick substrates, which makes transient flux measurements for significantly shorter periods infeasible.

Owing to the above-mentioned difficulties, almost all of the experimental and the numerical studies focus on

the steady state description of DWC. Limited data are available in literature for transient processes, which are

often encountered in practice. Parin et al. [41] performed a quasi-transient experimental study of DWC on nano-

structured aluminum surfaces. Their temporal resolution was on ~80 s. Numerical studies [1], [28] indicate that

such coarse resolution might be inadequate to capture the variation of transients during DWC. Results from the

present study, which uses high-intensity spray cooling, achieve start-up times of only a few seconds. The present

dissertation provides the first experimental measurements of transient heat transfer during the startup of DWC

at relatively high heat fluxes.

Prior numerical studies of dropwise condensation have either applied the concepts of fractals [30], [34]

or tracked individual droplets (Lagrangian approach) [1], [28]. These approaches have been able to give

satisfactory drop size distribution and average heat flux during steady state, with modest computational costs.

However, their transient predictions for startup have not yet been experimentally verified. The experimental

work described in this study also provides the first experimental verification for some of the earlier numerical

works. As implemented in the literature, these approaches have a number of limitations. In particular, these

approaches have treated droplets as rigid hemispheres, neglected the histories of droplets (e.g., non-quasi-steady

temperature distributions in droplets that have recently merged), and neglected the complex hydrodynamics

Page 21: A Study on Heat Transfer in Dropwise Condensation

7

associated with droplet merging and sliding. Multi-scale VOF simulations, which utilize the numerical and

experimental findings from the present dissertation could potentially mitigate these challenges.

1.2 Prior Research in Dropwise Condensation

1.2.1 Major Advances in the History of Dropwise Condensation

Dropwise condensation was first investigated rigorously more than 80 years ago by Schmidt et al. [42]

which garnered significant attention from the scientific community as the heat transfer coefficients were

reported to be an order of magnitude higher than in filmwise condensation. Studies by Tanzola and Wiedman

[43], Kajanne [44], Brunt and Minken [9], Poll et al. [45], etc. have found significant efficiency improvements

in various heat transfer systems enabled with dropwise condensation. Due to its industrial importance, most of

the literature deals with water/steam as the working fluid. Some researchers have studied relatively high surface

tension organic fluids (e.g., ethylene glycol, propylene glycol, and nitrobenzene etc.) [4], [46], [47] and metals

(for e.g., mercury) [48]–[51] as well. As in the case of water and steam, the heat transfer coefficients were found

to be significantly greater than those found in filmwise condensation, for the above-mentioned fluids.

1.2.2 Heat Transfer Models of Dropwise Condensation

Micro-cine studies by McCormick and Westwater [52], [53], and Peterson and Westwater [54] looked

into the process of droplet nucleation and established that droplets grow at specific nucleation sites. Umur and

Page 22: A Study on Heat Transfer in Dropwise Condensation

8

Griffith [16], studied and developed theory for droplet growth due to condensation, and validated their work by

comparing it with observed data by McCormick and Baer [20].

Due to limitations of the instrumentation and also lack of awareness about the effect of non-condensable

gases on the overall heat transfer during condensation, early experimental heat transfer data was inconsistent.

Later experiments took great care to eliminate non-condensable gases and used more sensitive instrumentation,

yielding several mutually agreeing heat transfer datasets. Studies by Tanner et al. [55], Graham [25],

Wilmshurst and Rose [56], Stylianou and Rose [5], and Tanaka and Tsuruta [57] have experimentally

characterized the average heat flux of a surface at steady state, resulting in a simple empirical equation for steam

condensation given by Rose[13].

q"[kWm−2] = T0.8 [5

ΔT

kv+ 0.3 (

ΔT

kv)2] 1.5

Here, T is the vapor temperature in Celsius, and kv is the thermal conductivity of the vapor.

Le Fevre and Rose [14], provided a theoretical approach in which heat flux through the base of a droplet

q"(r) of radius r is integrated over the entire drop size distribution over the condensing surface to predict the

overall average heat flux through the surface. Their theory has been corroborated and adjusted over the years

in the experimental studies by Rose [58] and Rose et al. [59]. The steady state distribution functions used in

such an approach have been verified by experimental works of Graham [25], and Tanasawa and Ochiai [26],

Numerical works of Glicksman and Hunt [28], and Burnside and Hadi [1], theoretical works of Rose and

Glicksman [2], and LeFevre [17], and others. Recently Kim and Kim [60] have formulated a theoretical drop

size distribution model for superhydrophobic surfaces (contact angle θ > 120°) and experimentally

corroborated it. The present work proposes to utilize such an approach to integrate q"(r) over the smaller size-

Page 23: A Study on Heat Transfer in Dropwise Condensation

9

scale portion of an evolving drop size distribution to predict average subgrid-scale (SGS) heat flux through a

surface. The SGS heat flux result can then be used for closure in a multi-scale simulation

Stylianou and Rose [5], Griffith and Lee [61], and others [62], [63] looked into the effect of the

condensing surface thermal conductivity on the overall heat transfer and found the effect to be negligible.

However, Tanner et al. [62], Hannemann and Mikic [40], and Wilkins and Bromley [63] reported that the effect

is significant. Later studies by Tsuruta et al. [64], [65] have seemed to confirm the importance of substrate

thermal conductivity. Analysis of experimental data in this project incorporates treatment for the constriction

resistance due to thermal conductivity of the condensing surface, which is found to be a significant effect.

A major challenge in the industrial application of dropwise condensation is developing methods to

prolong it in harsh industrial environment. Researchers used promoters like PTFE (e.g., Wilmshurst and Rose

[56]) or monolayer promoters like dioctadecyl disulphide (e.g., Le Fevre and Rose [14]). The interest in

dropwise condensation subsided in the 1980s as researchers found it difficult to find industrially reliable

promoters. This field of study was revived again in the 1990s with methods identified to promote dropwise

condensation [66]. Newer superhydrophobic and more durable surface treatments [67]–[75], which may last

longer in relevant environments, have led to renewed interest in dropwise condensation.

Recent advances in surface treatment technology have brought low adhesion, nano-textured

superhydrophobic surfaces into existence [76]–[82]. These recent developments have led many researchers to

focus their attention toward understanding the fundamental nature of droplet dynamics and droplet-surface

interactions [83]–[85]. The present work also aims to characterize droplet dynamics with heat transfer in greater

detail.

Page 24: A Study on Heat Transfer in Dropwise Condensation

10

1.2.3 Computational Approaches to Modeling Dropwise Condensation

Due to the experimental challenges, no measurements of the transient start-up of high-heat-flux

dropwise condensation have been reported, to the author’s knowledge. Some computational studies have been

performed to simulate this process. Glicksman and Hunt [28] performed a numerical study in which they tracked

the growth and coalescence of each droplet, treating them as individual liquid hemispheres. This approach can

be termed Lagrangian in the sense that it tracks evolution of individual parcels (droplets) rather than time or

space-averaged quantities (Eulerian approach). They modeled heat transfer assuming each droplet was at a

quasi-steady temperature distribution, and that heat flux did not depend on the history of individual droplets.

Moreover, coalescence of droplets was modeled as an instantaneous process and the intermediate complex

shapes of coalescing droplets, which could affect the time varying heat transfer rate, were neglected.

A similar study was conducted by Mei et al. [30]. As part of another study by Mei et al. [86], they

developed an analytical drop distribution model which was based on the concept of fractals and validated it

with a droplet tracking numerical simulation. They had a much greater number of initial nucleating droplets

which they tracked (around 50,000), compared with Glicksman and Hunt (around 1,000). They compared the

data with the experimental results of Tanasawa and Ochiai [26] and found good agreement. They conducted

numerical experiments with monodispersed (uniform size) and polydispersed initial droplets. They found that

the final steady state drop size distribution was insensitive to the initial distribution. This observation makes

their simulation technique suitable only for steady state studies. Their final predicted drop size distribution did

not match well at the extremities, that is, the smallest and the largest drop dimensions differed from

experimental observations.

Page 25: A Study on Heat Transfer in Dropwise Condensation

11

Using the concept of fractals, Wu et al. [34] developed a drop size distribution model. Their model was

developed by randomly populating square grids and iteratively superimposing them over each other to predict

the drop size distribution. Initially a coarse grid (cell length = r) was used and random cells were chosen to

represent drops according to a fixed χ. Then the grid was refined and the available area was repopulated with

the same value of χ. The grid was refined again and the process is repeated. This method can be continued for

several iterations depending on computing resources. For the finest grid, they predicted heat flux using a mean

heat transfer coefficient hr through the base of a droplet of radius r (model of Rose [22]) as the boundary

condition for the condensing surface. A steady state energy equation was solved to get the overall heat transfer

through the surface. Heat transfer coefficients over the surface were averaged to get the mean heat transfer

coefficient h of the surface.

Because the drop size spectrum is so large (6 orders of magnitude apart [13]), Wu et al. employed a

two-scale approach. In their approach, a much finer grid over was evaluated for a domain whose dimensions

were equal to a single cell on the coarser grid. The results of the finer grid were superimposed over the coarser

grid. This approach is sensitive to the chosen value of ��. To solve the transient heat transfer problem, Wu et al.

also developed a transient model for ��, which assumed quasi-steady behavior for each time step. This method

shares similarities with the multi-scale simulations proposed here, but relies on quasi-steady assumptions. For

example, their method does noes not account for droplet hydrodynamics.

While these prior Lagrangian numerical simulations have provided useful insights into the DWC

process, they share some common limitations. They do not account for the hydrodynamics of large droplets,

treating them instead as point particles. Additionally, quasi-steady temperature and size distributions are used

for closure, which may not apply in the transient startup of DWC.

Page 26: A Study on Heat Transfer in Dropwise Condensation

12

To enable more general predictive capabilities, the present investigation proposes a new multi-scale

transient simulation approach (Chapter 5). Coalescence of droplets is the primary mode of droplet growth for

water droplets ≥ 2 μm [15]. Therefore, it is important to study the hydrodynamics and the heat transfer process

during coalescence of droplets. Fidelity of the transient drop distribution and heat transfer models, to be

developed as part of the present work, can be improved with a proper understanding of the coalescence process

and its effect on the overall heat transfer on the surface. Cooper et al. [87], Sellier and Trelluyer [88], Wang et

al. [89], and Sikarwar et al. [90], etc. have looked into the coalescence process in detail, both numerically and

experimentally. However, these foregoing numerical and experimental studies have focused purely on the

hydrodynamics and do not consider heat transfer. Figure 1.2 shows the comparison between the droplet covered

surface of a droplet tracking numerical study by Burnside and Hadi [1], and another surface from the proposed

multi-scale numerical simulation approach (preliminary results). It can be clearly seen on the condenser surface

of the test case that several droplets are in the process of sliding and coalescing. These mechanisms allow larger

droplets to sweep the surface around them clean and initiate fresh nucleation sites, thereby affecting the overall

heat transfer on the surface.

Page 27: A Study on Heat Transfer in Dropwise Condensation

13

Figure 1.2 Comparison between droplet covered surfaces a) From the numerical study of

Burnside and Hadi [1]; b) One of the proposed multi-scale simulation test surfaces.

1.2.4 Goals of the Present research and Dissertation

The aim of this PhD dissertation research is to address the following fundamental questions about the

phenomena involved in dropwise condensation through both experimental and computational approaches.

I. Develop a Framework for Multi-Scale Simulations of Dropwise Condensation

One of the primary objectives of this work is to develop a framework for multi-scale simulation of

dropwise condensation. Such multi-scale simulations should capture both microscale heat transfer and large-

scale dynamics of dropwise condensation simultaneously with acceptable computational costs. Smaller droplets

are not very mobile, but account for a major portion of surface heat transfer. Therefore, a relatively simple

subgrid scale modeling approach that accurately captures the size distribution and heat transfer of small

Page 28: A Study on Heat Transfer in Dropwise Condensation

14

droplets, but neglects hydrodynamics, could be successful. Larger droplets account for a small portion of the

overall heat transfer directly (less than 10% of the total), but clear surface area during coalescence or sliding,

allowing small droplets to nucleate. Without accurately predicting such hydrodynamics, the rate of small droplet

nucleation and resulting intense heat fluxes would be incorrect. Thus, a successful simulation approach should

capture the hydrodynamics of such larger droplets. To achieve these simulation goals, the following steps are

undertaken:

a) Detailed Analysis of Heat Transfer in Individual Droplets

Models for dropwise condensation have generally assumed a statistical distribution of droplet sizes and

integrated heat transfer over the droplet size spectrum, considering droplet curvature effects on saturation

temperature, conduction thermal resistance, and interfacial resistance. Most earlier studies have assumed a

constant heat transfer factor (f = O(1)) to account for the relative contributions from conduction and interfacial

effects to total thermal resistance. However, f varies with droplet Biot number (Bi = ℎ𝑖𝑟

𝑘), where ℎ𝑖 is the

interfacial heat transfer coefficient of the droplet, r is the base radius, and k is the thermal conductivity and

contact angle (θ). Formulations for f with broad ranges of applicability are not currently available. In this

component of this dissertation, finite element simulations are performed to determine f and corresponding

numerical uncertainties for 0.0001 ≤ Bi ≤ 1000 and 10° ≤ θ ≤ 170°. This spans the active droplet size range

considered in most droplet condensation studies (e.g., for water condensing at Patm on a surface 10 K below the

ambient temperature, active droplets have 0.0005 < Bi < 300). A single explicit correlation is developed for f

and validated with available published data. A Volume of Fluid (VOF) two-phase model for capturing the

interfacial resistance is also developed to enable fully resolved simulations of condensing droplet heat transfer.

Page 29: A Study on Heat Transfer in Dropwise Condensation

15

This study will be used to validate the drop distribution and the average heat flux data from the proposed

experiments.

b) Characterizing Droplet Interactions During Coalescence

Prior models of dropwise condensation have generally described coalescence as an instantaneous event.

However, coalescence and recovery of a quasi-steady droplet temperature profile requires a finite time during

which the droplet heat transfer rate varies. A detailed description of droplet coalescence may be needed to

accurately model the overall dropwise condensation process. Information about condensing droplet heat transfer

during this process has not yet been reported. This portion of the dissertation employs simulations to show that

the transient heat transfer during coalescence (with water as the working fluid) can be fully characterized by a

decaying exponential function with a time constant 𝜏, radius of the smaller droplet 𝑟1, and Ratio (Rt) of the radii

of two coalescing droplets. Volume-of-Fluid (VOF) simulations are performed to determine the relationship

between 𝜏, 𝑟1 , and Rt (1 μm ≤ 𝑟1 ≤ 25 μm; 1 ≤ Rt ≤ 4). For water at Patm this spans the range of droplet sizes

through which most of the heat transfer occurs on a surface (~80%). An explicit correlation for 𝜏 is developed,

which can facilitate the analysis of transient heat transfer in dropwise condensation. The approach developed

in this study can be applied to resolve grid-scale hydrodynamics of larger droplets.

c) Multi-Scale Simulation of Dropwise Condensation

Previous computational studies of dropwise condensation have applied droplet tracking Lagrangian or

fractal-based approaches, in which the drop distribution pattern is assumed to be similar for all magnification

scales. In such studies, droplets are treated as quasi-steady rigid spheres and the droplet history effects are

neglected. Moreover, in these models, the process of coalescence has been assumed instantaneous. Therefore,

such studies may incorrectly predict the overall time-averaged heat transfer rate or have limited generality.

Page 30: A Study on Heat Transfer in Dropwise Condensation

16

Direct interface resolving approaches (e.g., VOF flow simulations) could directly predict condensing droplet

dynamics and heat transfer, but resolving the full droplet size spectrum during condensation would be

prohibitively expensive. In this study a multi-scale modeling approach is proposed that directly resolves large-

scale droplet dynamics with interface capturing flow simulations (e.g., VOF) and couples this with a subgrid

scale model for the evolving heat flux through smaller subgrid scale droplets between grid-scale droplets.

II. Experimental measurement of Transient Heat Transfer for Dropwise Condensation

The experimental thrust of this Ph.D. dissertation research seeks to collect the first experimental data

for early-transient heat transfer in high-flux dropwise condensation. While a number of experimental

investigations have been conducted to measure steady state dropwise condensation heat transfer rates, the

transient process has only been theoretically described through analytic and computational models. Although

some studies on breath figures [15], [91], [92] have considered DWC startup, they were performed at low heat

fluxes and significant non-condensable mass transfer resistances.

The transient startup process could last for a significant portion of the operating cycle of devices that

operate in a cyclic fashion, such as refrigerators or air-conditioners. To enable measurements during the early

transient startup process, the experimental setup will has a thin test section with low conduction resistance and

time scale, be exposed to a pure vapor (water is chosen as the working fluid due to its industrial and scientific

importance) environment on one side, and be cooled with a rapid supply of high-flow-rate coolant (spray

cooling) on the other side. Transient heat transfer data and high-speed visual data of the evolving drops will be

collected. These experiments are designed to enable assessment and closure of the proposed multi-scale

simulation approach.

Page 31: A Study on Heat Transfer in Dropwise Condensation

17

Chapter 2

Detailed Characterization of Individual

Droplet Heat Transfer1

1 Chapter 2 is adapted from the 2017 paper by Adhikari et al. [36], published in the International Journal of Heat and Mass

Transfer

Page 32: A Study on Heat Transfer in Dropwise Condensation

18

In dropwise condensation, heat from the condensation process transfers through the liquid-vapor

interface (interfacial resistance 𝑅int′′ ) and then conducts through the droplet to the cooled wall (Rcond) (Figure

1.1). Internal circulation and convection within droplets is generally assumed negligible [93]. Interfacial heat

transfer resistance is usually significant in dropwise condensation because the conduction length from the

interface to the wall approaches zero at the contact line. Therefore, interfacial thermal resistance is dominant in

the near-wall region. The relative overall contributions of interfacial and conduction resistances in droplet

condensation heat transfer depend on the Biot number (𝐵𝑖 =ℎ𝑖𝑟

𝑘𝑙) and contact angle 𝜃 [94]. Average heat flux

through the base of the droplet is given by the following relation:

𝑞" =

Δ𝑇

𝑓(Bi, 𝜃)𝑟𝑘𝑙

+12ℎ𝑖

2.1

Here, 𝑞" is the average heat flux through the base of the droplet, Δ𝑇 is the temperature difference

between ambient vapor temperature and the base, 𝑟 is the radius of curvature for the droplet, and 𝑘𝑙 is the liquid-

phase thermal conductivity. 𝑓(Bi, 𝜃) is a scaling factor for the conduction contribution to overall droplet thermal

resistance. 𝑓 can be considered as an inverse fin efficiency for the droplet on the cooled wall, and should

approach 0 for an infinitely conductive droplet (limBi→0 𝑓 = 0). Most prior studies assumed a constant factor f

= O(1) [1], [23], [28]. This assumption may be suitable for steady-state analyses where the distribution of drop

sizes is constant. However, an effective average f value would still be specific to the operating conditions. For

cases where the droplet size distribution evolves over time (e.g., dropwise condensation startup), the variation

of f as the droplet size distribution develops must be considered. Several studies have been performed to

determine formulae for f(Bi, θ). Sadhal and Martin [95] and Kim and Kim [60] have proposed analytical models

Page 33: A Study on Heat Transfer in Dropwise Condensation

19

for hydrophilic surfaces (θ < 90°). Recently, Miljkovic et al. [96] and Chavan et al. [35] have reported numerical

and experimental results for hydrophobic surfaces (θ > 90°). Generally valid models for f(Bi, θ) would facilitate

dropwise condensation analyses, but are not currently available.

In this chapter, finite element heat transfer analyses are performed for 2958 cases to develop a single

broadly applicable droplet condensation heat transfer formulation for 0.0001 ≤ Bi ≤ 1000 and 10° ≤ θ ≤ 170°.

Discretization errors are evaluated to determine uncertainty bounds for Richardson-extrapolated f factors.

Results are validated against prior findings from Sadhal and Martin [95] and Chavan et al. [35]. Summarized

data and simulation results from this study are publicly available at [97].

2.1 Modeling Approach

In this study, finite element analyses are performed for heat transfer in sessile droplets using COMSOL

MultiPhysics® (v. 5.2 [98]). In typical dropwise condensation conditions, more than 90% of the heat is

transferred through droplets less than 200 𝜇𝑚 in diameter [2]. Droplets in this size range are sufficiently small,

such that Marangoni circulation and other convection effects can be neglected [57]. Therefore, conduction is

the predominant mode of heat transfer. As droplets grow, the internal temperature distribution varies. However,

the conduction-front propagation velocity (𝛼𝑙

𝑟) is generally much greater than the droplet growth rate (

𝑑𝑟

𝑑𝑡);

therefore, the heat transfer process can be modeled as quasi-steady. For example, for a water droplet with 1 𝜇𝑚

diameter and a 90° contact angle on a surface at 99.15° C (𝑇𝑠𝑎𝑡 = 100.15° 𝐶) the conduction front velocity is

𝛼𝑙

𝑟~1.6 × 10−1 m s−1. Applying a standard droplet condensation heat transfer model (e.g., from [23] with

constant f = 0.25), the predicted growth velocity would only be 𝑑𝑟

𝑑𝑡= 8.4 × 10−5 m s−1. For similar conditions

Page 34: A Study on Heat Transfer in Dropwise Condensation

20

a 1 mm droplet will have 𝛼𝑙

𝑟= 1.6 × 10−4 m s−1 and

𝑑𝑟

𝑑𝑡= 6 × 10−7 m s−1. Based on these arguments, the in-

droplet condensation heat transfer process reduces to conduction in a solid medium, with the following

equation:

∇2𝑇 = 0 2.2

For small droplets, as is the focus here, surface tension forces dominate gravity. Therefore, the droplets

can be modeled as truncated spheres [35], and evaluated with 2-D axisymmetric domains (Figure 2.1). A fixed

temperature boundary condition is applied on the base of the axisymmetric droplet domain (1 K below the

saturation temperature), representing the cooled condenser surface temperature. A convection boundary

condition is applied to the liquid-vapor interface to account for interfacial resistance, with the following heat

transfer coefficient [99]:

ℎ𝑖 = (2𝜎

2 − 𝜎) ∗ (

ℎ𝑙𝑣2

𝑇𝑣𝑣𝑙) ∗ (

𝑀

2𝜋𝑅𝑇𝑣)1/2

2.3

Here, ℎ𝑖 is the interfacial heat transfer coefficient, 𝜎 is the accommodation constant, ℎ𝑙𝑣 is the enthalpy

of vaporization, 𝑇𝑣 is the ambient temperature, 𝑀 is the molecular mass of the working fluid, and R is the

specific gas constant. The saturation temperature of the vapor at the ambient pressure (𝑇𝑠𝑎𝑡) is substituted for

𝑇𝑣 in Eqn. 2.3.

The domain was meshed with nearly uniform size triangular elements (Figure 2.1). The governing

equation was discretized with second order accurate elements and solved to a relative residual of 10−14. The

Page 35: A Study on Heat Transfer in Dropwise Condensation

21

total droplet heat transfer rates obtained from the FEA were used to determine 𝑓(𝐵𝑖, 𝜃), using the relation given

in Eqn. 2.1, for varying Biot number and contact angle values.

Figure 2.1 Representative mesh sizes: (a) coarse 𝚫𝟑 = 𝟐. 𝟐 × 𝟏𝟎

𝟐𝐦𝐦; (b) fine 𝚫𝟐 =𝟏. 𝟐 × 𝟏𝟎−𝟐𝐦𝐦; (c) finest 𝚫𝟏 = 𝟔. 𝟎 × 𝟏𝟎

−𝟑𝐦𝐦

2.2 Simulation Studies

A parametric study for droplet condensation heat transfer was performed. The contact angle (𝜃) was

varied to span both the hydrophilic and hydrophobic regimes (10° – 170°) in 10° increments.Droplet

conductivity (𝑘𝑙) was varied with fixed radius (r = 1 mm) and interfacial resistance (ℎ𝑖=10000 W m-2 K-1) to

evaluate 174 Bi values from 0.0001 to 1000 for each contact angle. For water at atmospheric pressure, this Bi

range spans droplet radii from 310 nm to 3.1 mm. In total, 2958 cases were studied.

Grid sensitivity studies were performed for each case with at least three meshes to extrapolate

converged values of 𝑓(Bi, 𝜃) and corresponding uncertainties following the method of Celik et al. [100]. The

technique is briefly described below.

Page 36: A Study on Heat Transfer in Dropwise Condensation

22

Meshes (Figure 2.1) with average element dimensions: Δ3 = 2.2 × 10−2 mm, Δ2 = 1.2 × 10

−2 mm

and Δ1 = 6.0 × 10−3 mm are employed. Subscript 3 refers to the coarsest mesh and subscript 1 refers to the

finest mesh. The steps are as follows:

𝜀32 = 𝑓3 − 𝑓2; 𝜀21 = 𝑓2 − 𝑓1; 𝑠 = 𝑠𝑔𝑛 (

𝜀32𝜀21) 2.4

𝑅32 =

Δ3Δ2; 𝑅21 =

Δ2Δ1

2.5

𝑞(𝑝) = ln(𝑅21𝑝− 𝑠

𝑅32𝑝− 𝑠

) ; 𝑝(𝑞) =| ln (|

𝜀32𝜀21|) + 𝑞|

ln (𝑅21) 2.6

𝑓𝑒𝑥𝑡 = (𝑅21

𝑝𝑓1 − 𝑓2)/(𝑅21

𝑝− 1) 2.7

𝑒𝑎21 = |

𝑓1 − 𝑓2𝑓1

| ; GCI =1.25𝑒𝑎

21

𝑅21𝑝− 1

2.8

𝑢𝑛𝑐 = 𝑓𝑒𝑥𝑡 × GCI 2.9

The 𝜀 values are the changes in f (the heat transfer scaling factor) between meshes. R values are the

mesh refinement ratios. p is the empirical convergence rate, defined implicitly in terms of parameter q. 𝑓𝑒𝑥𝑡 is

the extrapolated value of f for Δ → 0 (infinitely fine mesh). 𝑒𝑎21 is the relative error between the two finest

meshes. GCI is the grid convergence index, a relative uncertainty estimate for 𝑓𝑒𝑥𝑡. The absolute uncertainty in

𝑓𝑒𝑥𝑡 (unc) is determined from Eqn. 2.9.

All studied cases were found to be monotonically converging (𝑠 = 1) and therefore the grid

convergence index (GCI) uncertainty estimate for 𝑓𝑒𝑥𝑡 was reported (<3% uncertainty in all cases). Summarized

data and simulation results from this study are publicly available at [97].

Page 37: A Study on Heat Transfer in Dropwise Condensation

23

Figure 2.2 shows the variation of 𝑓 with (Bi-1) for different contact angles (50° to 130°). Linear trends

of f with Bi-1 were found to continue for Bi = 10-1 – 10-4, and are not included in Figure 2.2 so that sharper

variations for large Bi are apparent. Similar trends of increasing f with Bi-1 for 𝜃 < 50° and decreasing f with

Bi-1 𝜃 > 130°, were found. Corresponding very low and high contact angle curves are therefore omitted from

Figure 2.2 for presentation clarity. Negative f values were found for hydrophobic droplets with low Bi values.

Such droplets have relatively low internal conduction resistances and greater surface area for condensation than

for the baseline θ = 90° droplet assumed in Eqn. 2.1. Thus, the negative f value yields a lower overall thermal

resistance than for interfacial resistance alone on a θ = 90° droplet.

Figure 2.2 𝒇(𝑩𝒊, 𝜽) vs. Bi-1 for different contact angles.

Page 38: A Study on Heat Transfer in Dropwise Condensation

24

2.3 Results and Discussion

2.3.1 Proposed Correlation

Simulation data (0.0001 ≤ Bi ≤ 1000, 𝜃 = 50° − 130°) for 𝑓(Bi, 𝜃) were fit to the following analytic

correlation using non-linear regression. The range of f is spread over 4 orders of magnitude. Therefore, data

points were weighted by 𝑓−2 in the fitting procedure to offset skew toward high magnitude values. The

correlation is given in Eqn. 2.10 and the coefficients are provided in Table 2.1. In Eqn. 2.10, the contact angle

𝜃 is in degrees.

𝑓(𝐵𝑖, 𝜃) = (𝐴 + 𝐵 ∗ 𝐵𝑖−1) + (𝐶 + 𝐷 ∗ 𝐵𝑖−1) ∗ 𝜃

+ [𝐸 ∗ log10(𝐵𝑖−1) + 𝐹 ∗ {log10(𝐵𝑖

−1)}2 + 𝐺 ∗ {log10(𝐵𝑖−1)}3]

2.10

Coefficient Value

(Uncertainty, 95% confidence interval)

A 0.2160 (±0.0015)

B 0.7278 (±0.0015)

C 0.001465 (±0.000013)

D -0.008086 (±0.000017)

E 0.1012 (±0.0007)

F -0.01378 (±0.00036)

G -0.007361 (±0.00013)

Table 2.1 Coefficients for the correlation given in Eqn. 2.10.

The absolute average deviation (AAD) for the data range (0.0001 ≤ Bi ≤ 1000, 50° ≤ θ ≤ 130°) is 7.4%.

Figure 2.3 compares the absolute values of f obtained from the correlation with conduction simulation results.

The correlation predicts 91% of cases with | f | < 1 within 25% of simulation values. All cases with | f | > 1 are

predicted within 25% of simulation values.

Page 39: A Study on Heat Transfer in Dropwise Condensation

25

Because of the wide range of f values, it was difficult to fit a simple analytic correlation over 10° ≤ θ ≤

170°. Therefore, a Matlab® [101] function for 𝑓(Bi, 𝜃) is provided [97] that directly interpolates simulation

results to provide precise values overall the full parametric space (0.0001 ≤ Bi ≤ 1000, 10° ≤ θ ≤ 170°).

Figure 2.3 Comparison of absolute values of f predicted with proposed correlation and

conduction simulations.

2.3.2 Validation of Results

For hydrophilic conditions the proposed correlation for f (Eqn. 2.10) was validated with data from

Sadhall and Martin [95]. Droplet Nusselt number Nu =𝑞

Δ𝑇𝑟𝑘 was compared. 𝑞 is the total heat transferred

through the base of a droplet. The AAD from the study, for the whole range of Bi and 𝜃, is 3.4%. Figure 2.4

Page 40: A Study on Heat Transfer in Dropwise Condensation

26

compares the variation of Nu with Bi. Numerical uncertainties are not presented with the simulation data points

in the figure because they are too small to be visible (< 10−4% ).

Figure 2.4 Nu vs. Bi for different contact angles (50° to 80°).

The only available data for f in hydrophobic droplets was reported by Chavan et al. [35]. They have

provided three correlations for 10-1 < Bi < 105. This span of Bi does not cover the typical full size range of

active droplets condensing on a surface. For example, for water condensing at 𝑃𝑎𝑡𝑚 on a surface 10 K below

the ambient temperature, active droplets have 0.0005 < Bi < 300 (0.0001 ≤ Bi ≤ 1000 evaluated in the present

Page 41: A Study on Heat Transfer in Dropwise Condensation

27

study). Good agreement between the proposed correlation (Eqn. 2.10) and the correlations of Chavan et al. [35]

are found in the overlapping data region (0.1 < 𝐵𝑖 < 1000, 90° < 𝜃 < 130°), with AAD = 10.93 %. The

comparison is provided in table format [97].

As there are no other published data to validate hydrophobic droplet condensation heat transfer, five

representative hydrophobic cases were also evaluated in interThermalPhaseChangeFoam [102], a volume-of-

fluid (VOF) based CFD solver for phase-change heat transfer processes. VOF is an interface capturing

technique, which tracks the volume fraction of the liquid phase in each mesh cell with a scalar 𝜙 ∈ [0, 1]. Eqn.

2.11 is the advection equation for 𝜙, where 𝑢𝑖 is the velocity vector and �� is the volumetric phase fraction

source due to condensation phase change. A single set of governing equations is solved for both the liquid and

the vapor phases. Fluid material properties are weighted by Eqn. 2.12.

𝜕𝜙

𝜕𝑡+

𝜕

𝜕𝑥𝑖(𝑢𝑖𝜙) = �� 2.11

𝛽 = 𝜙𝛽𝐿 + (1 − 𝜙)𝛽𝑉 𝛽 ∈ [𝜌, 𝜇, 𝑘] 2.12

To reduce the VOF computation time for the sessile, quasi-steady droplets considered here, the

momentum and phase fraction advection (Eqn. 2.11) equations are not evaluated (velocity field 𝑢 = 0, phase-

fraction field 𝜙 constrant). The steady state VOF thermal transport equation (Eqn. 2.13) is solved for a water

droplet (𝑟 = 1 𝑚𝑚) with different contact angles (90° to 130°). ��𝑝𝑐 is the volumetric condensation phase-

change heat source term applied on the interface. Interfacial resistance is applied as a distributed source term

on the diffuse interface region (~4 mesh cells thick). Volume generation is disabled to model the quasi-steady

conduction problem.

Page 42: A Study on Heat Transfer in Dropwise Condensation

28

𝜕

𝜕𝑥𝑖𝜕𝑥𝑖(𝑘𝑇) = ��𝑝𝑐 2.13

The VOF domain is axisymmetric and 4 mm× 4 mm in height and width. Meshes with average

element dimensions: Δ3 = 2.5 × 10−2 mm, Δ2 = 1.25 × 10

−2 mm and Δ1 = 6.25 × 10−3 mm are employed.

The base of the droplet is maintained at a temperature 2 K below the saturation temperature (𝑇𝑠𝑎𝑡 = 373.15° 𝐾).

The heat transfer in the vapor region is not included for comparison, and only the heat transfer through the base

of the droplet is employed for comparison with the COMSOL® (droplet-only) data. The results from

interThermalPhaseChangeFoam are also Richardson-extrapolated following the procedure from Eqns. 2.4 -

2.9.

Table 2.2 compares the average heat flux 𝑞", obtained using interThermalPhaseChangeFoam (VOF),

COMSOL conduction simulations, and the proposed correlation (Eqn. 2.10). Maximum error is less than 1.4

%. This close agreement between results from two different software packages with different simulation

methodologies further confirms the accuracy of reported hydrophobic droplet heat transfer data.

Biot No.

(Bi)

Contact

Angle (𝜽)

q” (W m-2)

VOF

q” (W m-2)

COMSOL

q” (W m-2)

Correlation

1.98 × 10−3 90° 8.87 × 105 ± 4300 8.75 × 105 ± 3.15 8.75 × 105

5.88 × 102 100° 5.43 × 103 ± 0.50 5.40 × 103 ± 700 5.57 × 103

2.40 110° 3.75E × 105 ± 6.0 3.72 × 105 ± 8600 3.82 × 105

0.10 120° 1.50 × 106 ± 7.0 1.47 × 106 ± 27000 1.43 × 106

3.68 × 10−1 130° 1.32 × 106 ± 3200 1.30 × 106 ± 38000 1.28 × 106

Table 2.2 Average heat flux through the base of droplets with representative hydrophobic Bi and 𝜃 values,

computed with VOF solver, COMSOL conduction studies, and proposed correlation.

Page 43: A Study on Heat Transfer in Dropwise Condensation

29

2.3.3 Internal temperature Distribution in Droplets

Temperature distribution trends inside droplets were found to depend more strongly on Bi than θ. As

Bi reduces, the conduction resistance also reduces relative to the interfacial resistance. Thus, as Bi reduces, the

internal droplet temperature distribution becomes more uniform. For higher Bi, internal conduction resistance

is significant and condensation heat transfer primarily occurs in a thin region near the contact line. Figure 2.5

presents the temperature distribution inside a droplet with varying Bi (𝜃 = 90°, droplet geometry and conditions

described in Section 2.2). It is evident from Figure 2.5 that for Bi < 0.2, the droplet is effectively isothermal.

This effect can also be observed in Figure 2.6, which presents the local heat flux at a given angle ranging from

𝛽 = 0° at the contact line to 𝛽 = 90° at the top of the droplet. For Bi = 0.2, condensation heat flux only varies

by 19% over the surface of the droplet. In contrast, for Bi = 1000 (high Bi), 90% of the heat transfer occurs in

the 20° of the droplet surface closest to the contact line.

Figure 2.5 Non-dimensional temperature (

𝑻−𝑻𝒘𝒂𝒍𝒍

𝑻𝒔𝒂𝒕−𝑻𝒘𝒂𝒍𝒍) distribution inside a droplet with varying

Bi. a) Bi = 1000, b) Bi = 2 and c) Bi = 0.2.

Page 44: A Study on Heat Transfer in Dropwise Condensation

30

Figure 2.6 Heat flux along the condensing surface of the droplet with 𝜽 = 𝟗𝟎° for different Bi.

𝜷 is the angle subtended by any point on the droplet surface and the horizontal (inset).

Trends of temperature distribution inside droplets were not found to depend strongly on the contact

angle. As can be observed in Figure 2.7, qualitatively similar temperature distributions are found for θ = 40° -

140° for a droplet at Bi = 1000.

Page 45: A Study on Heat Transfer in Dropwise Condensation

31

Figure 2.7 Non-dimensional temperature distribution inside droplets with different contact

angles for Bi = 1000. a) 𝛉 = 𝟒𝟎°, b) 𝛉 = 𝟗𝟎° and c) 𝛉 = 𝟏𝟒𝟎°.

2.3.4 Implications of varying Conduction Resistance Factor in Dropwise Condensation

Many prior dropwise condensation studies have assumed a constant value of conduction factor f.

However, this approximation can yield incorrect heat transfer predictions for cases with time-varying droplet

size distributions. As an example, consider the growth of a single water droplet (𝜃 = 90°, 𝑟 = 100 μm), on a

surface maintained 10 K below the saturation temperature (𝑇𝑠𝑎𝑡 = 373.15° 𝐾). The quasi-steady heat transfer

through a growing droplet is given as [103]:

𝑞 = 𝜌ℎ𝑙𝑣2𝜋𝑟2𝑑𝑟

𝑑𝑡 2.14

From Eqn. 2.14 and Eqn. 2.1, the growth rate of a droplet is:

𝑑𝑟

𝑑𝑡 =

Δ𝑇

2𝜌ℎ𝑙𝑣×

1

𝑓𝑟𝑘+

12ℎ𝑖

2.15

Page 46: A Study on Heat Transfer in Dropwise Condensation

32

Eqn. 2.15 is integrated for 1.5 s to predict the growth of the droplet. This analysis is performed

assuming both a constant value of f = 0.25 [28] and the proposed correlation for f (Eqn. 2.10), which varies

with the instantaneous droplet radius.

This droplet growth case is also evaluated using the interface-capturing volume-of-fluid (VOF) based

solver, interThermalPhaseChangeFoam to provide a high-resolution reference case. The domain is

axisymmetric and 400 μm × 400 μm in height and width. A fine mesh with an average element dimension Δ =

0.625 μm is employed. The momentum and phase fraction advection equations were solved here, unlike in the

quasi-steady cases described in the previous section.

Assuming a constant value of f yields a 23% relative error in the droplet growth with respect to the

VOF simulation (Δrf = 0.25 = 71 μm, ΔrVOF = 92 μm). Based on the trend at t = 1.5 s (Figure 2.8), this deviation

would continue to grow if the analysis were continued to a later time. However, the proposed correlation for f

results in only a 3% error. This analysis also indicates the accuracy of the proposed correlation, and its

applicability to unsteady droplet condensation cases.

Page 47: A Study on Heat Transfer in Dropwise Condensation

33

Figure 2.8 Comparison of droplet growth between OpenFoam VOF simulation, theoretical relation

with f obtained using the correlation and theoretical relation with f = 0.25.

2.4 Conclusion

In this study, finite element analyses were performed for condensation heat transfer in sessile droplets

accounting for internal conduction and interfacial resistance. Parametric studies were performed over a wide

range of Biot numbers (0.0001 ≤ Bi ≤ 1000) and contact angles (10° ≤ 𝜃 ≤ 170°) to determine conduction

thermal resistance factors (f) and corresponding uncertainties. Summarized data and simulation results from

Page 48: A Study on Heat Transfer in Dropwise Condensation

34

this study are publicly available at [97]. A new correlation was proposed for 𝑓(𝐵𝑖, 𝜃) that is valid over the

studied range of conditions. This model was validated against an earlier hydrophilic model, only applicable for

part of this range. Independent VOF simulations were performed to validate hydrophobic predictions.

An analysis was performed to highlight the importance of accounting for the variation of f with different

size droplets. As shown in Figure 2.8, assuming a fixed value of f can lead to large errors in droplet growth

predictions. The proposed, broadly applicable, correlation for f can thus facilitate future dropwise condensation

analyses and research.

It is important to note that the condensing surface was smooth for these studies and 𝜃 > 120° has not

yet been observed for smooth surfaces. For 𝜃 > 120° surface roughness needs to be accounted for, i.e., Biot

number for the substrate also becomes important for rough surfaces.

Page 49: A Study on Heat Transfer in Dropwise Condensation

35

Chapter 3

Characterizing Droplet Interactions During

Coalescence2

2 Chapter 3 is adapted from the 2018 paper by Adhikari et al. [146], published in the International Journal of Heat and Mass

Transfer

Page 50: A Study on Heat Transfer in Dropwise Condensation

36

Beysens and Knobler [15] described three regimes of dropwise condensation with a 100% relative

humidity nitrogen-steam mixture flowing over a cooled surface (vapor temperature 𝑇𝑉 = 23°C; surface

temperature 𝑇𝑆 = 21°C). Initially, in the first regime, for 𝑡 ≤ 1 s, a surface is covered by a homogenous pattern

of small droplets with both 𝑟 and 𝑎 smaller than 2 μm (here 𝑡 is time, 𝑟 is the radius of a droplet and 𝑎 is the

distance between two droplets). The droplets primarily grow by direct condensation [104]. In the second regime,

for 1 ≤ 𝑡 ≤ 300 s and 2 < 𝑟 ≤ 300 μm, the droplets primarily grow due to coalescence. In the third regime,

𝑡 > 300 s, the droplet size distribution on the surface reaches a statistically steady state condition, with

equilibrium rates of droplet removal, production, and growth. Their finding suggests that coalescence is the

dominant mode of growth for ambient pressure water droplets with r > 2 μm. Following the size distribution

model of Graham and Griffith [23] such droplets would account for 60-70% of the surface coverage during

dropwise condensation, indicating the importance of understanding the coalescence process.

A number of recent studies have investigated the details of droplet coalescence during dropwise

condensation. Chu et al. [105] have studied the effect of surface wettability and contact angle hysteresis on

droplet coalescence. Sprittles et al. [106], [107] have developed a finite element framework, which considers

formation and then gradual disappearance of an internal interface, to study the coalescence of two merging

droplets. Their numerical framework was also used by Enright et al. [108] to model coalescence induced

jumping of droplets on ultra-low adhesion surfaces. Enright et al. [108] also validated their findings

experimentally. Chen et al. [109] have also conducted experiments on superhydrophobic surfaces with

hierarchical microscale/nanoscale roughness to study coalescence induced jumping of droplets.

While many such investigations have advanced understanding of droplet coalescence hydrodynamics

[87], [88], [113]–[117], [105]–[112], limited information is available on heat transfer during coalescence.

Numerical simulations have been conducted of the overall dropwise condensation process, tracking growth of

Page 51: A Study on Heat Transfer in Dropwise Condensation

37

individual (ideal hemispherical) droplets in a Lagrangian sense [1], [30]. In these studies, the heat transfer rate

of individual droplets has been modeled in a quasi-steady function (independent of droplet history), with heat

transfer rates depending on current droplet radius, fluid properties, and operating conditions (surface

temperature and saturation temperature) [112]. Such studies have modeled coalescence as an instantaneous

event [1], [28], [30]. Typically, in these models, when two parent droplets overlap at the end of a simulation

time-step, they are replaced with a single larger hemispherical child droplet (volume conserved) with its center

at the volume-weighted mean of the parent droplet centers. At the next simulation time-step, the heat transfer

rate of the larger droplet corresponds to the quasi-steady value for its size. However, the actual droplet

coalescence proceeds over a finite time interval, and the resulting child droplet requires a period of time to

recover the quasi-steady temperature profile and heat transfer rate. Additionally, these models do not predict

the unsteady hydrodynamics of coalescence. Actual droplets can have complex time varying shapes during

coalescence, and the heat transfer rate depends on the shape of the droplet base [118]. Moreover, coalescing

droplets oscillate and sweep the area around them, repeatedly re-initializing nucleation on the surrounding

surface, and potentially increasing the overall heat transfer rate.

It is not yet known what role coalescing droplet transient heat transfer and hydrodynamics have on the

overall dropwise condensation process. Figure 3.1 shows a representative time series of the coalescence process

with two droplets of equal radii (25 μm), obtained using the interface tracking (Volume of Fluid – VOF)

simulation approach method presented later in this paper. This study seeks to provide initial quantitative data

on droplet coalescence heat transfer to inform improved models of dropwise condensation.

Page 52: A Study on Heat Transfer in Dropwise Condensation

38

Figure 3.1 Coalescing droplet mid-plane temperature profiles. Both parent droplet radii are 𝟐𝟓 𝛍𝐦.

The maximum velocity observed during coalescence is 𝐔𝐦𝐚𝐱 ~ 𝟑 𝐦 𝐬−𝟏.

Considering the simplified coalescence assumption of prior modeling studies of dropwise

condensation, the present study seeks to:

1. Perform interface resolving simulations of droplet coalescence to identify heat transfer trends and

timescales

2. Determine the extent to which the instantaneous coalescence assumption under-predicts

condensing droplet heat transfer rates

3. Formulate an analytic model for time-dependent heat transfer after droplet coalescence, which

could be incorporated into Lagrangian simulations of dropwise condensation to improve heat

transfer predictions

4. Estimate the degree of heat transfer enhancement due to repeated clearing of droplets from the

surrounding wall area by oscillating coalescing droplets

Page 53: A Study on Heat Transfer in Dropwise Condensation

39

To achieve these goals, this study performs Volume-of-Fluid (VOF) simulations of coalescence of pairs

of water droplets in saturated steam environment at 100 kPa on a previously characterized surface (hydrophobic

silane). The VOF approach is validated using hydrodynamic data from the literature [95] and a set of

conduction-only Finite Element (FEM) heat transfer simulations (based on the approach given by Adhikari et

al.[36]). A parametric study is performed for varying droplet radii (𝑟 ∈ [1, 100] μm – most of the heat ~80%

is transferred through this range of droplet sizes [2]) and varying ratios of the merging droplet radii (Rt ∈ [1, 4]).

It is shown that transient condensation heat transfer during coalescence can represented by an exponential curve

with time constant 𝜏. A simple correlation is proposed for 𝜏 for the range of conditions in this parametric study.

A representative coalescence case is simulated with a subgrid scale model to estimate the heat transfer

enhancement due to an oscillating droplet repeatedly clearing surrounding microscopic droplets.

3.1 Modeling Approach

3.1.1 Governing equations

This study employs the Volume of Fluid (VOF) interface tracking formulation, as implemented in

OpenFOAM® [119]. VOF is an interface capturing technique in which the liquid volume fraction in each

computational cell is represented by a scalar field 𝜙 ∈ [0,1] (𝜙 = 0 all vapor in cell, 𝜙 = 1 all liquid in cell),

and a single set of governing mass, momentum, and thermal energy transport equations is solved for both

phases. The phase fraction field is advected by the solved velocity field (Eqn. 3.1). Fluid material properties at

each point are determined as a 𝜙-weighted average of liquid and vapor values (Eqn. 3.2, 3.3).

Page 54: A Study on Heat Transfer in Dropwise Condensation

40

𝜕𝜙

𝜕𝑡+

𝜕

𝜕𝑥𝑖(𝑢𝑖𝜙) = ��𝑝𝑐 3.1

𝛽 = 𝜙𝛽𝐿 + (1 − 𝜙)𝛽𝑉 𝛽 ∈ [𝜌, 𝜇, 𝑘] 3.2

𝑐𝑝 =

𝜙𝑐𝑝,𝐿𝜌𝐿 + (1 − 𝜙)𝑐𝑝,𝑉𝜌𝑉𝜙𝜌𝐿 + (1 − 𝜙)𝜌𝑉

3.3

Here 𝑢𝑖 is the velocity vector, ��𝑝𝑐 is the volumetric phase fraction source due to condensation phase

change, 𝛽 represents the fluid property in a cell and the subscripts 𝐿 and 𝑉 represent the liquid phase and the

vapor phase, respectively. 𝜌 is the density, 𝜇 is the viscosity, 𝑘 is the thermal conductivity, and 𝑐𝑝 is the mass

weighted specific heat capacity in a cell [102], [120], [121]. The fluid properties are assumed constant for the

vapor and liquid phases.

The present simulations focus on the period between initiation of coalescence between two parent

droplets and the recovery of a quasi-steady temperature profile in the child droplet. During this period, the

increase in liquid volume in a droplet due to condensation is assumed to be small. Additionally, the momentum

transfer from the condensing vapor to the droplets is relatively low. Therefore, the process can be modeled

assuming incompressible flow (Eqn. 3.4, ��𝑝𝑐 = 0) and by using the incompressible Navier-Stokes equation

(Eqn. 3.5).

𝜕𝑢𝑖𝜕𝑥𝑖

= 0 3.4

𝜕(𝜌𝑢𝑖)

𝜕𝑡+

𝜕

𝜕𝑥𝑗(𝜌𝑢𝑖𝑢𝑗) −

𝜕

𝜕𝑥𝑗[𝜇 (

𝜕𝑢𝑖𝜕𝑥𝑗

+𝜕𝑢𝑗

𝜕𝑥𝑖)] = −

𝜕𝑝

𝜕𝑥𝑖+ 𝜌𝑔𝑖 + 𝑓𝑠,𝑖 3.5

Page 55: A Study on Heat Transfer in Dropwise Condensation

41

Here, 𝑔𝑖 is the gravity vector and 𝑓𝑠,𝑖 is the surface tension force (applied on a volumetric basis,

described in Section 3.1.2).

The justification for the incompressible flow assumption is as follows. The characteristic velocities (𝑉𝑐)

due to droplet coalescence are much greater than the droplet growth rate (𝑑𝑟

𝑑𝑡). 𝑉𝑐 is defined as the distance

moved by the larger of the two droplets divided by the time it takes for the oscillations to die out. The droplet

front velocity due to condensation phase change is 𝑑𝑟

𝑑𝑡 (Eqns. 3.6, Eqn. 3.7) [122]. Based on simulation results

described later, for a water droplet with 1 μm diameter on a surface at 99.15° C (𝑇𝑠𝑎𝑡 = 100.15° C)

𝑉𝑐 ~ 8.3 m s−1. The predicted droplet growth rate due to condensation is:

𝑑𝑟

𝑑𝑡 ~ 8.4 × 10−5 m s−1. For similar

conditions, a 1 mm droplet will have 𝑉𝑐 ≈ 2.4 × 10−3 m s−1 and

𝑑𝑟

𝑑𝑡= 6 × 10−7 m s−1.

𝑑𝑟

𝑑𝑡 =

Δ𝑇

2𝜌ℎ𝐿𝑉×

1

𝑓𝑟𝑘+

12ℎ𝑖

3.6

ℎ𝑖 = (2��

2 − ��) ∗ (

ℎ𝐿𝑉2

𝑇𝑉𝑣𝐿) ∗ (

𝑀

2𝜋𝑅𝑇𝑉)1/2

3.7

Here, Δ𝑇 is the temperature difference between the cooled surface and the vapor, k is the thermal

conductivity of the liquid, and 𝑓 = 𝑓(Bi, θ) is the coefficient for conduction resistance [36], ℎ𝑖 is the interfacial

heat transfer coefficient, �� (0.03 [122]) is the accommodation constant, ℎ𝐿𝑉 is the enthalpy of vaporization, 𝑇𝑉

is the ambient temperature, 𝑀 is the molecular mass of the working fluid, and R is the specific gas constant.

The saturation temperature of the vapor at the ambient pressure (𝑇𝑠𝑎𝑡) is substituted for 𝑇𝑉 in Eqn. 3.7.

Page 56: A Study on Heat Transfer in Dropwise Condensation

42

Similarly, the relative significance of the momentum imparted by steam vapor condensing into the

interface can be assessed by comparing the vapor stagnation pressure (1

2𝜌𝑉 [

𝑑𝑟

𝑑𝑡(1 + 𝜌𝐿/𝜌𝑉)]

2) to the capillary

pressure (2𝜎 𝑟⁄ ). For these cases, the vapor stagnation pressure (r = 1 μm, Δp = 5 × 10-3 Pa; r = 1 mm, Δp = 3

× 10-7 Pa) is negligible compared with the capillary pressure (r = 1 μm, Δp = 1.2 × 105 Pa; r = 1 mm, Δp = 1.2

× 102 Pa). Based on these arguments, the growth of the interface due to phase change and condensation of vapor

into the interface do not affect the hydrodynamics of the problem, justifying the incompressible flow

assumption.

Eqn. 3.8 is the thermal energy equation, where 𝑘 is the thermal conductivity and ��𝑝𝑐 is the volumetric

phase change heat source. Further details are provided in Section 3.1.3.

𝜕(𝜌𝐶𝑝𝑇)

𝜕𝑡+

𝜕

𝜕𝑥𝑖(𝜌𝐶𝑝𝑢𝑖𝑇) =

𝜕

𝜕𝑥𝑖[(𝐶𝑝𝑘)

𝜕𝑇

𝜕𝑥𝑖] − ��𝑝𝑐 3.8

3.1.2 Surface Tension Model

For the droplets size range of the present study Marangoni effects can be neglected [57], [123].

Therefore, a constant surface tension value is assumed (𝜎 = 0.056 N m−1). Surface tension forces are

calculated in each cell using the continuum surface tension force formulation (Eqn. 3.9) developed by Brackbill

et al. [124]. Resulting momentum fluxes through cell faces are applied as an explicit source term in the pressure

correction process.

Page 57: A Study on Heat Transfer in Dropwise Condensation

43

𝑓𝑠,𝑖 = 𝜎𝜅∇𝜙; 𝜅 = −(∇ ∙ ��); �� =

∇𝜙

|∇𝜙| 3.9

Here, 𝜎 is the surface tension between the liquid and the vapor phases, 𝜅 is the curvature at the interface,

and �� is the unit normal into the vapor phase. In the OpenFOAM VOF implementation, the interface is spread

over a few cells; therefore, ∇𝜙 is effectively smoothed delta function which applies the force over the diffuse

interface.

3.1.3 Phase Change Model

The phase change model is based on the approach of Rattner and Garimella [120]. The volumetric heat

source ��𝑝𝑐 is applied on interface cells (Eqn. 3.8).

��𝑝𝑐 =

ℎ𝑖𝐴Δ𝑇

𝑉𝑐𝑒𝑙𝑙; 𝐴 = |∇ϕ| ∙ Vcell 3.10

Here, ℎ𝑖 is the interfacial heat transfer coefficient, 𝐴 is an approximation of the area of the interface

area contained in a cell, Δ𝑇 is the difference between the saturation temperature and the local temperature (Δ𝑇 =

𝑇𝑠𝑎𝑡 − 𝑇), and 𝑉𝑐𝑒𝑙𝑙 is the volume of the cell. ��𝑝𝑐 is solved implicitly as part of the energy equation (Eqn. 3.8).

The foregoing heat transfer model has been validated by Adhikari et al. [36].

Page 58: A Study on Heat Transfer in Dropwise Condensation

44

3.1.4 Dynamic Contact Angle Model

A static contact angle model is insufficient to predict the process of coalescence because the three-

phase contact line moves during the process. In general, the advancing contact angle 𝜃𝐴, is greater than the

static contact angle θ0, which is greater than the receding contact angle 𝜃𝑅 (Figure 3.2). To model the

coalescence accurately, the dynamic contact angle model of Yokoi et al. (Eqn. 3.11) [125] is used:

𝜃𝐷 =

{

min [θ0 + (

Ca

ka)

13, θma] 𝑖𝑓 Ca ≥ 0

max [θ0 + (Ca

kr)

13, θmr] 𝑖𝑓 Ca < 0

3.11

Here, 𝜃0 is the static contact angle, 𝜃𝐷 is the dynamic contact angle, Ca =μUcl

𝜎 is the local capillary

number, Ucl is the velocity of the moving contact line which is positive in the direction of the droplet outward

normal vector (��) on the condensing surface. ka and kr are liquid-surface pair dependent constants, 𝜃𝑚𝑎 is the

limit for advancing contact angle, and 𝜃𝑚𝑟 is the limit for receding contact angle. In the present study parameters

for water-silicon (with hydrophobic silane grafted on it) are used, as listed by Yokoi et al. [125]. The surface

parameters are as follows: 𝜃𝑚𝑎 = 114°, 𝜃𝑚𝑟 = 52°, 𝜃 = 90°; k𝑎 = 9 × 10−9; and kr = 9 × 10

−8. The model,

as implemented here, is validated with experimental data from [125] in Section 4.2.2.

Page 59: A Study on Heat Transfer in Dropwise Condensation

45

Figure 3.2 Advancing and receding contact angles measured using volume changing technique

(liquid injection or removal from the top of the droplet). a) Increasing volume of the drop. b)

Decreasing volume of the drop.

3.1.5 Solution Algorithm

The OpenFOAM® finite volume formulation is applied to discretize the governing equations, and the

solution algorithm is summarized as a flow chart in Figure 3.3. Expanded discussion of the phase change

simulation approach is provided in the study by Nabil and Rattner [102].

Page 60: A Study on Heat Transfer in Dropwise Condensation

46

Figure 3.3 Phase change volume-of-fluid solver algorithm flow chart.

Page 61: A Study on Heat Transfer in Dropwise Condensation

47

3.2 Simulation Studies

3.2.1 Studied Domain and Case setup

Coalescence of two droplets is symmetric along one center plane. Therefore, the domain only includes

half of the droplets and their surrounding region. Figure 3.4 shows the representative domain with labeled

boundaries.

Figure 3.4 Representative domain for the VOF coalescence study. The face showing the droplets is

the symmetry plane; the bottom surface, which is not visible in the figure, is the condensing

surface; and the remaining 4 faces are open atmosphere.

Page 62: A Study on Heat Transfer in Dropwise Condensation

48

Here, 𝐿 = 6 × 𝑟1 + 2 × 𝑅𝑡 × 𝑟1 and 𝑊 = 𝐻 = 2 × 𝑟1 + 𝑅𝑡 × 𝑟1, 𝑟1 is the radius of the smaller droplet,

and Rt ≥ 1 is the ratio of the two merging droplets. Table 3.1 shows the boundary conditions at the respective

boundaries.

Boundary Boundary condition

Symmetry

∇T ∙ �� = 0

∇𝑃 ∙ �� = 0

𝑈 ∙ �� = 0; ∇𝑈 ∙ �� = 0

∇𝜙 ∙ �� = 0

Atmosphere

∇T ∙ �� = 0

𝑃 = 𝑃𝑎𝑡𝑚

∇𝑈 ∙ �� = 0

∇𝜙 ∙ �� = 0

Wall (condensing surface)

𝑇 = (𝑇𝑠𝑎𝑡 − 1 K)

∇𝑃 ∙ �� = 0

𝑈 = 0

�� ∙ ��𝑤 = cos(𝜃𝐷)

Table 3.1 Boundary conditions for the VOF simulation domain

Here ��𝑤 is the unit normal from the wall pointing into the domain, �� is the unit normal from the liquid

phase to the vapor phase, and �� is the unit normal to the patch. For wall ��𝑤 = ��. The dynamic contact angle

model is applied as a boundary condition for the phase fraction term 𝜙. While the no-slip boundary condition

is applied at the wall, in this finite volume formulation, the mean velocity in cells adjacent to the wall can be

Page 63: A Study on Heat Transfer in Dropwise Condensation

49

non-zero. This introduces enough numerical slip to eliminate the stress singularity at the three-phase contact

line [126].

For the studied case of steam condensing at atmospheric pressure, more than 95% of the heat transfer

occurs through droplets with r < 100 μm [14], [23] – which is the radius of the largest droplet in this parametric

study. Surface tension dominates gravity for such small droplets; therefore, the droplets can be initialized as

truncated spheres [35].

Figure 3.5 Steady state temperature distribution on the droplet surfaces before merging.

For the studied range of droplet sizes (1 μm < r < 100 μm) conduction is the primary mode of heat

transfer [57]. Conduction is a diffusion process [127], and therefore, the thermal front takes a finite amount of

time to propagate from the base of the droplet to the liquid vapor interface. An order of magnitude estimate of

the conduction front propagation velocity is (𝛼𝑙

𝑟). This velocity is generally much greater than the droplet

growth rate (Section 3.1.1). For example, for a water droplet with 1 μm diameter on a surface at 99.15° C

(𝑇𝑠𝑎𝑡 = 100.15° 𝐶) the conduction front velocity is 𝛼𝑙

𝑟~1.6 × 10−1 m s−1, whereas the droplet growth rate is

𝑑𝑟

𝑑𝑡 ~ 8.4 × 10−5 m s−1. Based on the arguments, a quasi-steady state internal temperature distribution can be

Page 64: A Study on Heat Transfer in Dropwise Condensation

50

assumed in droplets just before coalescence begins. Therefore, the droplets are initialized with a steady-state

temperature distribution (Figure 3.5), obtained by solving the steady-state energy equation (Eqn. 3.12).

𝜕

𝜕𝑥𝑖𝜕𝑥𝑖(𝑘𝑇) = ��𝑝𝑐 3.12

Page 65: A Study on Heat Transfer in Dropwise Condensation

51

3.2.2 Grid Resolution, Sensitivity Study and Dynamic Contact Angle Validation

The mesh resolution must be sufficient to resolve heat transfer and hydrodynamic effects. Figure 3.6

shows the representative droplet isosurfaces with VOF meshing, for the parametric study. The mesh is made

of cube cells, and extends similarly in the third dimension.

Figure 3.6 Representative droplet isosurfaces, indicating VOF mesh resolutions for the

droplets from the parametric study. a) For smaller droplet radius 𝒓𝟏 = 𝟏 𝛍𝐦 and radius ratio

Rt = 1. b) For 𝒓𝟏 = 𝟏 𝛍𝐦 and Rt = 2. c) For 𝒓𝟏 = 𝟏 𝛍𝐦 and Rt = 4.

Page 66: A Study on Heat Transfer in Dropwise Condensation

52

Grid resolution in the domain is defined using 𝑛𝑟 = (𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑐𝑒𝑙𝑙𝑠)/𝑟1. Conduction grid

independence studies are carried out for the largest and the smallest radii droplets (𝑟1 = 1 μm and 𝑟1 = 25 μm).

𝑛𝑟 is varied from 5 to 21, and the heat transfer rate (𝑄) through the base of the droplets is evaluated. For each

mesh in the grid sensitivity study, 𝑛𝑟 is increased by a factor of 1.6 (total number of cells increases by 4×). 𝑄

is plotted against 𝑛𝑟. In both the cases, it is observed that 𝑄 becomes asymptotic (Richardson extrapolated

values of Q are given in the supplemental data available publicly [128]) when 𝑛𝑟 ∈ [13,21] (Figure 3.7). The

uncertainties for Q in both the cases are less than 0.4%. Therefore, 𝑛𝑟 = 15 can be assumed the minimum

resolution for heat transfer in the parametric study (Figure 3.6-c). As an example, for a droplet of radius 1 μm

and 𝑛𝑟 = 15, average length of a mesh element Δ = 6.67 × 10−8 m.

Figure 3.7 Heat transfer grid independence study (Q vs 𝒏𝒓 where 𝒏𝒓 = # 𝒄𝒆𝒍𝒍𝒔/𝒓𝟏) for 𝒓𝟏 =𝟏 𝛍𝐦 and 𝒓𝟏 = 𝟐𝟓 𝛍𝐦.

To study hydrodynamic grid independence, and also to verify the selected dynamic contact angle

model, an axisymmetric droplet impact simulation is performed for the case of [125], [126]. In the simulation

Page 67: A Study on Heat Transfer in Dropwise Condensation

53

a 2.28 mm diameter spherical water droplet impacts a surface with a velocity of 1 m s−1. The surface properties

are given in Section 3.1.4, and the fluid properties are as follows: dynamic viscosity liquid 𝜇𝑙 =

1 × 10−3 kg m−1 s−1; dynamic viscosity air 𝜇𝑔 = 1.82 × 10−5kg m−1 s−1; 𝜎 = 0.072 N m−1. 𝑛𝑟 for the

axisymmetric study is varied from 19 to 55. Figure 3.8 presents the variation of average diameter D (square

root of the base area) of the impacting droplet as a function of time for different grid resolutions. Experimental

data [125] is also presented. It is observed that the average absolute deviation (AAD) for D from 𝑛𝑟 = 38 to

𝑛𝑟 = 55 is less than 3%. Since, in the parametric study, the droplets merge and oscillate together as a larger

droplet before settling, 𝑛𝑟 = 40 is selected as the minimum resolution for the parametric study (Figure 3.6-a),

which is a stricter requirement than imposed by the heat transfer process alone.

Figure 3.8 Colliding droplet contact patch diameter on surface (D) vs. time (t) for different

grid resolutions and experimental data [125].

Page 68: A Study on Heat Transfer in Dropwise Condensation

54

3.3 Validation of Results

Response time and spatial resolution requirements make it difficult to study the process of droplet

coalescence experimentally. For example, in the present study: 1 μm < 𝑟 < 100 μm, and the times of

coalescence are only 0.5 μs < 4 × |𝜏| < 1.2 ms. During the startup of dropwise condensation on a surface,

droplets could be even smaller. Owing to these challenges, there are no experimental or computational studies

available to validate the time-varying heat-transfer during coalescence. However, the steady state heat transfer

results from this simulation (just before and long after coalescence) can be compared with published steady-

state conduction heat transfer simulation data from Sadhal and Martin [95] to validate the heat transfer

formulation used in the study.

Steady condensation heat transfer rates for water droplets of r = 1 – 100 μm radii (the range of sizes

studied here) from this simulation study are compared with results from [95]. The average absolute deviation

(AD) from that study is less than 3%. The comparison is given in Figure 3.9, where the total heat transferred

through the base of the droplet (Q) is plotted against the Biot number (Bi =ℎ𝑖𝑟

𝑘) of the droplet. Here ℎ𝑖 is the

interfacial heat transfer coefficient, r is the radius, and k is the thermal conductivity of the droplet. Bi

corresponds directly to r in the simulation as hi and k are fixed.

Page 69: A Study on Heat Transfer in Dropwise Condensation

55

Figure 3.9 Comparison of the simulation data (heat transfer rate (Q) vs. Biot number (Bi)) with

that of Sadhal and Martin [95], and conduction-only finite element COMSOL simulations.

A finite element conduction heat transfer analysis is also performed for droplets of r = 1 – 100 μm to

further validate the simulation heat transfer formulation. For the studied range of droplet sizes Marangoni

circulation and other convection effects can be neglected [57]. Therefore, the finite element analysis reduces to

a simple conduction problem in truncated spheres. The droplets are evaluated with 2-D axisymmetric domains

(Figure 3.10). A fixed temperature boundary condition is applied on the base of the axisymmetric droplet

domain (1 K below the saturation temperature), representing the cooled condenser surface temperature. A

convection boundary condition is applied to the liquid-vapor interface to account for interfacial resistance hi.

The conduction studies were performed in COMSOL Multiphysics® (v. 5.2 [98]).

Page 70: A Study on Heat Transfer in Dropwise Condensation

56

Grid sensitivity study was performed for each finite element case where three meshes were utilized to

extrapolate the converged value of total heat transfer Q through the base of any droplet. The method is given

by Celik et al. [100], and is briefly described here.

Figure 3.10 Representative mesh sizes: (a) coarse 𝚫𝟑 = 𝟏. 𝟗𝟕 × 𝟏𝟎

−𝟗𝛍𝐦− 𝟏. 𝟗𝟗 × 𝟏𝟎−𝟓𝛍𝐦; (b)

fine 𝚫𝟐 = 𝟓. 𝟖𝟒 × 𝟏𝟎−𝟏𝟎𝛍𝐦− 𝟓. 𝟗𝟎 × 𝟏𝟎−𝟔𝛍𝐦; (c) finest 𝚫𝟏 = 𝟏. 𝟒𝟖 × 𝟏𝟎

−𝟏𝟎 𝛍𝐦 −𝟏. 𝟓𝟎 × 𝟏𝟎−𝟔𝛍𝐦.

Meshes (Figure 3.10) with average element dimensions: Δ3 = 1.97 × 10−9μm− 1.99 × 10−5μm,

Δ2 = 5.84 × 10−10μm− 5.90 × 10−6μm and Δ1 = 1.48 × 10

−10 μm − 1.50 × 10−6μm are employed.

Subscript 3 refers to the coarsest mesh and subscript 1 refers to the finest mesh. The steps are as follows:

𝜀32 = 𝑄3 − 𝑄2; 𝜀21 = 𝑄2 − 𝑄1; 𝑠 = 𝑠𝑔𝑛 (

𝜀32𝜀21) 3.13

𝑅32 =

Δ3Δ2; 𝑅21 =

Δ2Δ1

3.14

𝑞(𝑝) = ln(𝑅21𝑝− 𝑠

𝑅32𝑝− 𝑠

) ; 𝑝(𝑞) =| ln (|

𝜀32𝜀21|) + 𝑞|

ln (𝑅21) 3.15

𝑄𝑒𝑥𝑡 = (𝑅21

𝑝𝑄1 − 𝑄2)/(𝑅21

𝑝− 1) 3.16

Page 71: A Study on Heat Transfer in Dropwise Condensation

57

𝑒𝑎21 = |

𝑄1 − 𝑄2𝑄1

| ; GCI =1.25𝑒𝑎

21

𝑅21𝑝− 1

3.17

𝑢𝑛𝑐 = Qext × GCI 3.18

The 𝜀 values are the changes in Q between meshes. R values are the mesh refinement ratios. p is the

empirical convergence rate, defined implicitly in terms of parameter q. 𝑄𝑒𝑥𝑡 is the extrapolated value of Q for

Δ → 0 (infinitely fine mesh). 𝑒𝑎21 is the relative error between the two finest meshes. GCI is the grid convergence

index, a relative uncertainty estimate for 𝑄𝑒𝑥𝑡. The absolute uncertainty unc in 𝑄𝑒𝑥𝑡 is determined from Eqn.

3.18.

All studied cases were found to be monotonically converging (𝑠 = 1) and therefore the grid

convergence index (GCI) uncertainty estimate for 𝑄𝑒𝑥𝑡 was reported (<0.2% uncertainty in all cases). Figure

3.9 shows the comparison with simulation results.

The hydrodynamics (Figure 3.8) and steady-state heat transfer rates (Figure 3.7) are both separately

validated. This supports the validity of the simulation approach for predicting transient heat-transfer in

coalescing droplets.

3.3.1 Results and Discussion

Figure 3.1 shows a representative time series evolution of surface temperature profiles as two parent

droplets (r1 = 25 μm; Rt = 1) merge together to form a new child droplet. Heat transfer through the base of the

merging droplets is presented. Initially, the droplets have quasi-steady temperature profiles, which are disturbed

at the start of coalescence. When the two droplets come in contact, a liquid bridge forms (Figure 3.1, 0 μs).

Page 72: A Study on Heat Transfer in Dropwise Condensation

58

The bridge initially has concave curvature, resulting in a local low-pressure zone due to surface tension forces.

Liquid flows from the outer high pressure regions in the parent droplets toward the low pressure region in the

bridge [129] (0 – 20 μs). Due to contact angle hysteresis, the outer portion of the three-phase contact line does

not begin to move until well after the start of coalescence (~25 μs). At this point, the child droplet begins

oscillating on the condensing surface (30 – 200 μs). The oscillations gradually reduce in intensity due to viscous

dissipation within the droplet. The temperature profile within the droplet also decays toward a quasi-steady

profile (Figure 3.1, 500 μs). For these simulations of water condensing at atmospheric pressure, the oscillation

period and the heat transfer decay period are found to be approximately equal.

A decaying exponential function, with time constant 𝜏, can be used to represent the time varying

average heat flux through the child droplet. A parametric study is performed to characterize heat flux trends in

coalescing droplets over a range of sizes (r1) and radius ratios (Rt = r2/r1). Table 3.2 provides the simulation test

matrix and along with the values of 𝜏 obtained from the simulations. The working fluid is saturated water at

atmospheric pressure.

Droplet Radii [𝛍𝐦] Ratio (Rt = r2/r1) Time constant (𝝉) [𝛍𝐬]

r1 r2

1 1 1 0.129

1 2 2 0.357

1 4 4 0.650

6.25 6.25 1 5.80

6.25 12.5 2 16.1

6.25 25 4 28.9

25 25 1 65.9

25 50 2 170

25 100 4 287

Table 3.2 Simulation test matrix for the parametric study of coalescing droplets, with varying smaller

radius (r1) and the ratio between the radii of the merging droplets (Rt). Time constant 𝜏 is obtained

from the simulation results.

Page 73: A Study on Heat Transfer in Dropwise Condensation

59

The following sections characterize the process of coalescence and propose trends for the stabilization

time of coalescing water droplets at atmospheric pressure for the studied range of droplet sizes and surface

parameters. The impact of coalescence on condensation heat transfer is assessed considering direct heat transfer

through a child droplet (direct mechanism), and repeated clearing of the surrounding surface as the droplet

oscillates (indirect mechanism).

3.3.2 Characterizing Coalescence

In dropwise condensation, latent heat from the condensation process transfers through the liquid-vapor

interface (interfacial resistance 𝑅int′′ ) and then conducts through the droplet to the cooled wall (Rcond). Some

studies have argued that the overall single droplet heat transfer rate depends on the three-phase contact line

length [130]–[132], as most of the heat is transferred through the perimeter of the droplet, near the condenser

surface (Figure 3.5). This result applies for droplets with high Bi (Biot number Bi = ℎ𝑖𝑟

𝑘), for which conduction

resistance is much greater than the interfacial resistance. However, for cases where the conduction resistance is

comparable with the interfacial resistance (low Bi), significant conduction occurs through the entire base of the

droplet, and not just the contact-line region. Figure 3.11 shows the dependence of heat transfer on the droplet

radius. The working fluid for the figure is saturated water at atmospheric pressure. The calculations were based

on the heat transfer model of Adhikari et al. [36].

Page 74: A Study on Heat Transfer in Dropwise Condensation

60

Figure 3.11 Total heat transfer rate (Q) vs. droplet radius (r) for water condensation at atmospheric

pressure. The dashed lines indicate the rate of variation of heat transfer with droplet radius for

different size droplets.

The form of the curve in Figure 3.11, for the range r ∈ [3 × 10−10, 3 × 10−4] m (Q ∈

[1 × 10−12, 1] W), is approximately quadratic (i.e., Q ∝ r2). This suggests that for smaller droplets, heat

transfer is proportional to base area.

Page 75: A Study on Heat Transfer in Dropwise Condensation

61

Figure 3.12 compares the time variation of the heat transferred through two equal size coalescing

droplets with 𝑟1 = 𝑟2 = 6.25 μm (Bi ≈ 0.3) with the time variation of the base area of the merging droplets and

the perimeter length of the contact line. It is clear from the comparison that heat transfer rate is proportional to

base area during the coalescence of water droplets of this size range.

Page 76: A Study on Heat Transfer in Dropwise Condensation

62

Figure 3.12 . Comparison of a) condensing droplet heat transfer rate vs. time, b) base area vs.

time, and c) contact line (perimeter) length vs. time for two merging drops of radii 𝟔. 𝟐𝟓 𝛍𝐦

(𝐁𝐢 ≈ 𝟎. 𝟑).

A slight increase in the heat transfer is observed shortly after the beginning of coalescence. This occurs

because the total base area briefly increases while the liquid bridge grows, but the outer contact line sections

Page 77: A Study on Heat Transfer in Dropwise Condensation

63

are still pinned due to hysteresis. Once the hysteresis effect is overcome, the droplet base area and heat transfer

sharply decrease, and then oscillate in-sync.

A Finite Fourier Transform (FFT) can be applied to transient heat transfer traces to extract the principal

frequencies in the oscillating coalescence process. As an example, the principal frequency of the time varying

heat transfer through droplets (𝑅1 = 𝑅2 = 1 μm) is 1 × 105 Hz. The average heat transfer trend can be

approximated with a decaying exponential function (Figure 3.13), neglecting this oscillatory component.

Figure 3.13 Decaying exponential function superimposed over Q vs Time plot of two equal

merging drops of radii 𝟔. 𝟐𝟓 𝛍𝐦 each.

Figure 3.13 presents the best-fit decaying exponential function and the oscillatory time-varying heat

transfer through the base of two merging droplets (𝑟1 = 𝑟2 = 6.25 μm; time constant τ = 5.8 × 10−6 𝑠). This

Page 78: A Study on Heat Transfer in Dropwise Condensation

64

heat transfer decay rate could be incorporated into Lagrangian-type models of the overall dropwise

condensation process [13] to more accurately account for transient heat transfer after coalescence events, rather

than assuming a step change from the quasi-steady heat transfer rate of the two parent droplets to that of the

single child droplet.

3.3.3 Effect of Coalescence on Heat Transfer

If a dropwise condensation model assumes instantaneous merging and recovery of quasi-steady heat

transfer for a pair of droplets, it will under-predict the total heat transfer. The resulting heat transfer deficit is

due to a step-change assumption is:

Def = (

∫ 𝑎𝑒−𝑡𝜏

𝑡coalesce0

𝑑𝑡

∫ 𝑎𝑒−𝑡𝜏

𝑡coalesce0

𝑑𝑡+𝑐𝑡coalesce

) × 100 %. 3.19

Here, a is the difference in the steady state heat transfer between that of the two parent droplets and the

child droplet. c is the steady state heat transfer for the child droplet. The decaying heat transfer is integrated

from time t = 0 to t = tcoalesce – the time for the child droplet to settle to a steady heat transfer rate.

This heat transfer deficit effect is illustrated for two representative cases: (1) 𝑟1 = 1 μm, Rt = 1 amd

(2) 𝑟1 = 25 μm,Rt = 4. For cases 1 and 2, Def1 ~ 8% and Def2 ~ 5%. This deficit could become significant if

many droplets are coalescing on a surface.

Assuming instantaneous droplet coalescence thus causes a modest under-prediction of heat transfer by

neglecting the thermal stabilization time of the child droplet. This could be termed as the direct mechanism of

Page 79: A Study on Heat Transfer in Dropwise Condensation

65

heat transfer enhancement due to coalescence. However, merging droplet hydrodynamics can cause a second

indirect heat transfer enhancement mechanism. As merging droplets oscillate before stabilizing, they repeatedly

clear small droplets from surrounding surface area, and initiate periodic renucleation in this swept area. Local

condensation heat flux can be very high right after droplets are cleared from a region. Therefore, this indirect

heat transfer enhance mechanism may be even more significant than the direct mechanism.

Glicksman and Hunt [28] performed a numerical study of the dropwise condensation process from

initialization on a dry surface, yielding transient heat transfer trends. Data from their study is adopted here in

this paper to estimate the indirect heat transfer enhancement mechanism (curve fit to their data in Eqn. 3.20).

Their data are for saturated water at atmospheric pressure (Tsat = 100°C) with a condensing surface 0.28°C

below the Tsat. Nucleation site density was assumed to be 109 cm−2.

𝑞"𝑆𝐺𝑆(𝑡) = (8.379 × 105 W m−2) (

𝑡

1 s)−0.2356

3.20

Here, t represents the time since the condensing surface was initially dry. The transient heat flux trend

from [28] (𝑞"𝑆𝐺𝑆(𝑡)) is applied as a subgrid scale model on portions of the wall not covered by grid scale

droplets in the VOF simulation. This heat flux represents the heat transfer from condensing small droplets

around the studied grid-scale droplets. Here, 𝑞"𝑆𝐺𝑆(𝑡) is evaluated for every cell face on the wall boundary. A

local time (t) is evaluated for each face as the time since it was last covered by a grid-scale droplet. 𝑞"𝑆𝐺𝑆 is set

to zero when a wall face is covered by a grid-scale droplet. Thus, this model captures the effect of merging

droplets periodically sweeping the surface around them and re-initializing nucleation in the swept area.

Page 80: A Study on Heat Transfer in Dropwise Condensation

66

This subgrid scale model is evaluated for two cases with r1 = 25 μm;Rt = 1. In the first case, two

droplets coalesce and oscillate, as described in Section 3.3.1. The subgrid scale heat flux value on the wall faces

is initialized to 0. A second case represents the results of assuming instantaneous coalescence. Here, a single

stationary droplet (same total liquid volume as in case 1) is placed in the center of the domain with a quasi-

steady temperature distribution. The subgrid scale heat flux value on the wall is initialized to 0, except for the

area under the two parent droplets, which are set to t = 0 s. The heat flux in these cells evolves over time

according to Eqn. 3.20. The difference in heat transfer between the two cases (oscillating interface tracked

droplet coalescence case and the static case) serves as an estimate of the heat transfer underprediction by

assuming instantaneous coalescence. This resulting indirect heat transfer deficit is:

Defindirect = (

∫ 𝑞"𝑆𝐺𝑆_𝐷𝑦𝑛𝑎𝑚𝑖𝑐𝑡𝑐𝑜𝑎𝑙𝑒𝑠𝑐𝑒0

𝑑𝑡 − ∫ 𝑞"𝑆𝐺𝑆_𝑆𝑡𝑎𝑡𝑖𝑐𝑡𝑐𝑜𝑎𝑙𝑒𝑠𝑐𝑒0

𝑑𝑡

∫ 𝑞"𝑆𝐺𝑆_𝑆𝑡𝑎𝑡𝑖𝑐𝑡𝑐𝑜𝑎𝑙𝑒𝑠𝑐𝑒0

𝑑𝑡) × 100% 3.21

Figure 3.14 shows the different trends of subgrid scale heat transfer between the dynamic case and the

static case. Figure 3.14-a shows the dynamic case in which the oscillating child droplet repeatedly clears

sections of the wall, reinitializing nucleation and high heat fluxes. For the static case (Figure 3.14-b), however,

𝑞"𝑆𝐺𝑆(𝑡) is only non-zero on that exposed area that was previously under the parent droplets.

Page 81: A Study on Heat Transfer in Dropwise Condensation

67

Figure 3.14 Differences between subgrid scale heat transfer due to microscale condensing

droplets around two coalescing r = 25 μm droplets: a) the dynamic coalescence case; and b) the

static coalescence case. Top-down view snapshots of the droplet shapes, represented in black,

are presented. The colored area underneath the droplets shows the evolution of subgrid scale

heat flux due to small condensing droplets around the primary coalescing droplets

(𝐪"𝐒𝐆𝐒∗ (𝐭) =

𝐪"𝐒𝐆𝐒−𝐦𝐢𝐧(𝐪"𝐒𝐆𝐒(𝐭))

𝐦𝐚𝐱(𝐪"𝐒𝐆𝐒(𝐭))−𝐦𝐢𝐧(𝐪"𝐒𝐆𝐒(𝐭))).

Page 82: A Study on Heat Transfer in Dropwise Condensation

68

The heat transfer trends for the two cases are presented in Figure 3.15. After 350 μs, when the

oscillations in the droplet have almost died out, the heat transfer deficit approaches Defindirect ~ 11%, which

is greater than the direct deficit. Unlike the direct deficit, the magnitude of this indirect deficit is expected to

increase for larger oscillating droplets.

Figure 3.15 Cumulative heat transfer for an interface tracked dynamic coalescence case and a

static case. Initially, for both cases, both droplets have radius 𝒓 = 𝟐𝟓 𝛍𝐦.

Page 83: A Study on Heat Transfer in Dropwise Condensation

69

3.3.4 Simulation Data and Proposed Correlation for Droplet Coalescence Time Constant

Simulation data for τ(r1, Rt) from the parametric study cases listed in Table 3.2 were used to form a

correlation (Eqn. 3.22) for τ for coalescence of condensing water droplets at atmospheric pressure on the

studied hydrophobic silane surface. τ and r1 were non-dimensionalized (τ∗ = 𝜏𝜌𝜎2/𝜇3; r1∗ = 𝜎𝜌𝑟1/𝜇2) using

the simulation parameters 𝜌, 𝜎 and 𝜇 (for water at 1 atm), and a correlation was formed for τ∗(r1∗, Rt). Table

3.3 lists the constants given in Eqn. 3.22. Correlation τ∗ vs. r1∗ curves for varying Rt are compared with

simulation data (markers) in Figure 3.16.

τ∗ = a × r1∗b × Rtc 3.22

Constant Value

a 0.063

b 1.927

c 1.144

Table 3.3 Constants for the droplet coalescence time constant correlation (Eqn. 3.22)

The absolute average deviation (AAD) between the correlation and the data is 14%, and the maximum

error is 37%. This represents reasonable agreement considering the three order of magnitude range of τ* in the

study space.

Page 84: A Study on Heat Transfer in Dropwise Condensation

70

Figure 3.16 Dimensionless time constant (𝝉∗) vs. dimensionless smaller droplet radius (𝒓𝟏

∗ ) and

droplet radius ratio Rt. The curves are generated using the correlation given in Eqn. 3.22 and the

simulation data is represented by the markers.

Two additional test cases: (1) 𝑟1 = 3 μm, Rt = 1, and (2) 𝑟1 = 12.5 μm,Rt = 2, were simulated to

assess the correlation given in Eqn. 3.22. The comparison is provided in Table 3.4. The relative deviation for

both cases is less than 30%.

𝒓𝟏 [𝛍𝐦] Rt 𝛕 𝐒𝐢𝐦𝐮𝐥𝐚𝐭𝐢𝐨𝐧 [𝛍𝐬] 𝛕 𝐜𝐨𝐫𝐫𝐞𝐥𝐚𝐭𝐢𝐨𝐧 [𝛍𝐬] AD (%)

3 1 1.36 1.11 18

12.5 2 53.61 38.34 28

Table 3.4 𝝉 for representative test cases obtained from best-fit exponential curves to the experimental

data and the proposed correlation (Eqn. 3.22).

Page 85: A Study on Heat Transfer in Dropwise Condensation

71

3.4 Conclusions

In this study, a VOF formulation was proposed and validated for interface dynamics predictions

(droplet impact) and steady state droplet condensation heat transfer. This method was then applied to simulate

the process of water droplet coalescence during dropwise condensation for 9 cases over a range of radii (1 μm ≤

r ≤ 100 μm) and radius ratios (1 ≤ Rt ≤ 4). Results were used to determine the time constant for droplet

stabilization and transient heat transfer trends after dropwise condensation. The working fluid in this study was

saturated water at atmospheric pressure. The surface properties (hydrophobic silane) were adopted from the

study of Yokoi et al.[125]

Two mechanisms (Section 3.3.3) were identified for heat transfer enhancement due to droplet

coalescence. The direct mechanism is due to the gradual decay of heat transfer from the quasi-steady rate for

the two parent droplets to that of the larger child droplet. Instantaneous droplet coalescence models neglect this

thermal stabilization time for the child droplet, and therefore under-predict droplet heat transfer by 5-10%

during the period of coalescence. The indirect mechanism is due to the oscillating child droplet repeatedly

clearing small droplets from the surrounding wall area, reinitializing nucleation in those areas. For two

coalescing droplets of equal radii (r1 = 25 μm), heat transfer enhancement due to the indirect mechanism is

~11%. This indirect effect may be even greater in larger droplets that oscillate over larger areas and have more

momentum (longer oscillation periods).

Page 86: A Study on Heat Transfer in Dropwise Condensation

72

A correlation for τ(r1, Rt) was presented for the studied range of droplet sizes, fluid properties, and

surface conditions. Two test cases (r1 = 3 μm,Rt = 1 and r1 = 12.5 μm,Rt = 2) were simulated to test the

accuracy of the correlation, and satisfactory agreement was found.

Prior theoretical and numerical studies of dropwise condensation have generally assumed droplet

coalescence to be instantaneous, and that child droplets immediately recover quasi-steady internal temperature

profiles (no history effect) [28], [34]. The present study used a direct interface tracking simulation technique to

characterize the process of droplet coalescence and post-coalescence heat transfer trends. Findings suggest that

the assumption of instantaneous coalescence underestimates the total heat transfer (direct underprediction +

indirect underprediction due to the oscillating child droplet sweeping the area underneath) in the time shortly

after droplet coalescence by 15-20%. Results from this study provide an improved closure model for droplet

coalescence that can be incorporated into Lagrangian dropwise condensation models to improve overall heat

transfer predictions. Such improved dropwise condensation models may provide insights into the effects of

coalescence and oscillating droplet dynamics on overall surface heat transfer.

Future investigations are warranted to expand results for coalescing droplet dynamics and heat transfer

to a broader range of fluids and surfaces. In order to achieve that, a better surface description is needed. For

example, Biot number of the substrate, Bisub can be introduced to account for heat transfer on a rough surface.

Investigation of more complex coalescence dynamics is also needed, such as a cascade of coalescence events

initiated by merging of a single droplet pair. If these complex droplet coalescence processes are better

understood, it may be possible to enhance dropwise condensation heat transfer through surface engineering.

For example, if contact angle hysteresis can be reduced, the oscillation time and spatial span of coalescing

droplets may increase, amplifying the two heat transfer enhancement mechanisms identified in this study.

Page 87: A Study on Heat Transfer in Dropwise Condensation

73

Chapter 4

Heat Transfer Measurements of Fast

Transient Startup of Dropwise Condensation

Page 88: A Study on Heat Transfer in Dropwise Condensation

74

Dropwise condensation (DWC) was first studied rigorously over 80 years ago by Schmidt et al.[42],

and has received sustained research interest because it can yield heat transfer coefficients (HTC) an order of

magnitude higher than in filmwise condensation [34], [59], [103], [133]. Industrial and commercial applications

have been explored, including thermal management [8], [134], power generation [135], desalination [7], and

fogging on transparent surfaces [136]. Many experimental [14], [49], [103], [137], [138] and numerical [28],

[36], [60], [74], [86], [90] studies have been performed to characterize fundamental aspects of DWC, including

initial droplet formation [92], [93], droplet size distribution and density statistics [104], [138]–[140], and critical

sizes for droplet growth and departure [28], [141]. Numerical studies by Glicksman and Hunt [28], and Burnside

and Hadi [1] predict that instantaneous HTCs during the startup period of dropwise condensation could be much

greater than steady state values. Steady state description is not adequate for some applications, such as vapor

chambers for thermal management of devices with pulsed heat fluxes [142], or refrigeration cycles [143] that

operate for short periods of time to hold some temperature. Therefore, it is imperative to characterize transients

during DWC.

Both numerical and experimental studies face different sets of problems dealing with transient DWC.

Numerical studies in the past have primarily focused on droplet tracking [28], [144] and drop-size distribution

on a surface as fractals [33], [145]. These studies have treated droplets as rigid spheres with quasi-steady

temperature profiles. Coalescence events have generally been treated as being instantaneous. As such, both the

droplet temperature history and the complex hydrodynamics during coalescence have been neglected, which

may have significant effects on the heat transfer response of the condensing surface [146]. It is unclear whether

such modeling approaches are suitable for the startup process.

Most of the experiments in literature have employed temperature sensors in the substrate, under the

condensing surface, at varying depths[14], [26], [38]–[40]. Surface temperature can be extrapolated from such

Page 89: A Study on Heat Transfer in Dropwise Condensation

75

measurements using Fourier’s law. However, this approach may be difficult to adapt to transient heat transfer

measurements of DWC startup. During startup, the temperature profile through the substrate will be non-linear,

making instantaneous heat flux estimation difficult. Further, conventional DWC heat transfer test sections have

relatively thick substrates, which may incur conduction timescales comparable to that of the overall DWC

startup process. For example, in the seminal experimental study by Hannemann and Mikic [40], flux is

measured using thermocouples (spaced along th = 31.75 mm) in a copper rod, which is pressed against a

stainless-steel condenser plate. The conduction time scale τcond can be estimated as th2

αcopper, where αcopper is

the thermal diffusivity of copper. Here, 𝜏𝑐𝑜𝑛𝑑 ~ 𝑂(10) s, which makes transient flux measurements for

significantly shorter periods infeasible.

Owing to the above-mentioned difficulties, almost all of the experimental and the numerical studies

focus on the steady state description of DWC. Limited data are available in literature for transient processes,

which are often encountered in practice. For example, some researchers [90], [146] are working on interface

resolving numerical techniques, like Volume of Fluid (VOF), to overcome the drawbacks associated with

previously mentioned techniques. Since, the drop size distribution spans 6 orders of magnitude, the latter

approach cannot resolve all the droplets. Therefore, transient heat transfer data from experimental studies is

needed to supplement these numerical studies.

Parin et al.[41] performed a quasi-transient experimental study of DWC on nano-structured aluminum

surfaces. They utilized the inlet-outlet temperature difference of the coolant to calculate the heat flux. Surface

temperatures were calculated using Fourier’s law. They report HTCs as a function of time, but their temporal

resolution was only one reading per ~80 s. Numerical studies[1], [28] indicate that such coarse resolution might

Page 90: A Study on Heat Transfer in Dropwise Condensation

76

be inadequate to capture the variation of transients during DWC at industrially relevant heat fluxes. Results

from the present study, which uses high-intensity spray cooling, indicate start-up times of only a few seconds.

The objective of this investigation is to provide the first experimental measurements of transient heat

transfer during the startup of DWC at relatively high heat fluxes (250 – 350 kWm−2). A uniquely instrumented

test section and facility are designed to collect these transient measurements (c.f. Section 4.1.2). Time varying

surface temperature measurements are then used as boundary conditions for an inverse heat transfer code to

determine instantaneous heat fluxes and the HTCs. Steady-state portions heat data are compared with available

models from the literature and reasonable agreement is obtained, indicating validity of the experimental facility.

Transient data are then interrogated and reveal three distinct phases of the startup period, which emphasizes the

importance of this characterization for devices operating in transient mode.

The above mentioned three distinct phases characterize the startup in dropwise condensation and the

minimum HTC period in phase II emphasizes the importance of this characterization for devices operating in

transient mode.

4.1 Experimental Design and Setup

An experimental facility was developed to record transient space-averaged heat fluxes and high-speed

video of the DWC startup process. The overall facility is designed to closely reproduce the conditions assumed

in prior numerical models of the DWC startup process [28], [147]. Typically, these studies assume that an

initially dry surface is exposed to a step change in cooling, causing instantaneous startup of DWC on a face

exposed to a constant pressure pure steam atmosphere. In this experiment, condensation occurs on a thin

composite test plate (~1.5 mm aluminum core), which yields a short characteristic conduction time scale

Page 91: A Study on Heat Transfer in Dropwise Condensation

77

(τ ~ δ2/α = 58 ms). One side of the plate is exposed to pure condensing steam supplied from a boiler with

automated temperature control. The plate is resistively heated and dried between runs, and a solenoid valve can

be opened to rapidly deliver high intensity spray cooling on the back face to initiate DWC. Thin resistive

temperature devices (RTDs) printed on both faces provide high-frequency temperature measurements, which

are supplied to an inverse 1-D conduction model to determine transient heat fluxes and HTCs.

Conceptually, the test plate (c.f. Section 4.1.3) must be thin enough to limit transient conduction effects

and delay, but thick enough to yield an appreciable temperature difference between faces for accurate heat flux

and HTC measurements. Non-condensable gases can concentrate near condensing surfaces and significantly

lower the heat fluxes [55], [62], [148]. Therefore, the facility must also incorporate provisions to remove these

gases during operation (c.f. Section 4.1.1).

4.1.1 Apparatus

Page 92: A Study on Heat Transfer in Dropwise Condensation

78

Figure 4.1 Experimental setup to capture transient heat transfer data during DWC. a) Schematic

of the setup, b) Photograph of the full test stand, and c) photographs of major components: (1)

Steam generator; (2) Test section; (3) Test plate with resistance temperature detectors (RTDs)

printed on both sides and heater for drying between tests; (4) Full cone spray nozzle; (5) Primary

pressurized cooling water tank; (6) Auxiliary water collection tank; (7) Venting and vacuum

arrangement for steam; (8) Peripheral heater for the test plate; (9) constant current source for

RTDs and signal amplifier; (10) Data acquisition system (DAQ); (11) View-port for camera.

Page 93: A Study on Heat Transfer in Dropwise Condensation

79

The experimental facility is shown schematically in Figure 4.1-a, photograph of the full test setup is

shown in Figure 4.1-b, and photographs of key components are shown in Fig. 4.1-c. Provisions for high speed

photography of the condensation process are detailed in Section 4.1.2. The test plate (3, c.f. Section 4.1.3) acts

as a pressure barrier that separates the test section (2, c.f. Section 4.1.2) into a sealed steam side and an open

spray-cooling side. Distilled water is boiled in the steam generator (1), which has an electronic process

controller to maintain delivery pressure. Riser and downcomer tubes passively supply steam and return

condensate to the test section. Non-condensable gases are purged from the steam loop via the valve at point (7).

Between tests, the steam generator is heated while a dry-scroll vacuum pump extracts vapor and gas through

this valve. A degassing cycle is repeated several times until recorded steam temperatures and pressures in the

loop correspond to saturated states (i.e., for the pressure transducer and temperature sensor in the test chamber,

Pmeasured = Psat(Tmeasured). The steam side also has a vent valve, which is open to the atmosphere and continuously

exhausts during experimental runs to minimize build-up of noncondensables. Resistive heaters (8) are used to

dry the test plate between runs.

Filtered tap water coolant is stored in a primary tank (5) pressurized with compressed air. A solenoid

valve can be actuated to rapidly initiate intense spray cooling via a full cone spray nozzle directly behind the

test plate (4). The spray cooling HTC during the experimental run is 10 − 20 kW m−2 K−1 (c.f. Section 4.2.4).

The spent coolant is collected in an auxiliary tank (6) and pumped back to the primary tank (5) between tests.

Temperature measurements from the two sides of the test plate are acquired using a purpose built 4-

wire RTD amplifier board (9). These readings and signals from other instruments are digitized with a high-

speed data acquisition system (10, LabJack® T7), which can sample up to 100,000 signals per second.

Page 94: A Study on Heat Transfer in Dropwise Condensation

80

4.1.2 Test Section

Figure 4.2 Test Section: 3 – Test plate; 4 – Full cone spray nozzle; 2.1 – Coolant supply line;

2.2 – Coolant discharge port to the auxiliary tank; 2.3 – Condensate return port; 2.4 – Steam

inlet; 2.5 – Sight glass assembly; 2.6 – Port for venting, vacuum, and pressure transducer; 2.7 –

Temperature sensor port. (Dimensions in mm)

Figure 4.2 presents a detailed cross section view of the test section. (2.1) is the coolant supply line for

spray cooling, (2.2) is the discharge port to the auxiliary water tank (Figure 4.1 (6)), (2.3) is the return line for

steam, (2.4) is the steam inlet, (2.5) is the sight glass assembly for a view port, (2.6) is the port used in venting,

measuring pressure and pulling vacuum in the chamber, and (2.7) is the port for installing the calibrated

thermocouple. The main body of this chamber is constructed from stainless steel pipe and flanges. The test plate

(3) is sandwiched between two pipe stubs, and is sealed on its faces with custom annular cup-type silicone

gaskets. The right side of the test section is a pure steam and water environment (steam side). The left side of

the assembly is open to the atmosphere and cooling system (coolant side). The entire test section, steam inlet

and condensate outlet are insulated to minimize heat loss on the steam side. Additionally, a resistive heater (not

shown) dries the test plate between runs. The sight glass is also heated to remove condensation. Calibrated

Page 95: A Study on Heat Transfer in Dropwise Condensation

81

pressure measurements (± 1 kPa) on the steam chamber are used to determine instantaneous saturation

temperatures during DWC. A calibrated thermocouple (± 0.2 K) is also installed in the chamber for redundant

measurements of the vapor state and to detect the presence of non-condensable gases in the steam environment.

Figure 4.3 a) Evolution of steam saturation temperature 𝐓𝐄 with time; b) Evolution of steam

pressure 𝐏𝐄 with time.

Figure 4.3 shows both saturation temperature and pressure, as functions of time, during the experiment.

The saturation temperature (Figure 4.3-a) is derived directly from the pressure (Figure 4.3-b). The pressure

data is collected at a frequency (f) of 4000 Hz, and later down-sampled using a moving average filter (period

𝑛 = 100) to smooth out the data. The vent valve (Figure 4.2, (8)), as mentioned earlier, is kept open during the

entire run to limit buildup of non-condensable gases in the steam chamber. The pressure gauge (Figure 4.2

(2.6)) is calibrated with a thermocouple (Figure 4.2 (2.7)), which itself is calibrated (±0.2 K) with a high

accuracy RTD thermometer. The calibration process is as follows. The vent can be partially closed to different

degrees to maintain different steady state pressures in the test section as the steam is vented. Temperature

Page 96: A Study on Heat Transfer in Dropwise Condensation

82

readings from the calibrated thermocouple are recorded at different pressure points, which are later used to

calibrate the pressure gauge. Since this calibration process does not involve significant condensation, the effect

of non-condensable gases is negligible. Readings from the pressure gauge are used to calculate the steam

saturation temperature during condensation experiments as it responds much faster than the thermocouple.

Pressure and thermocouple readings from condensation tests are also compared with the saturation curve for

steam and the mean deviation is less than 0.3 K. Therefore, it is safe to conclude that the steam side is a nearly

pure steam environment.

Page 97: A Study on Heat Transfer in Dropwise Condensation

83

Figure 4.4 a) High speed camera arrangement; b) Camera view during the experimental run.

Figure 4.4-a shows the high-speed camera aimed directly at the condensing surface of the test plate,

and Figure 4.4-b shows its view. The camera is rigidly mounted to the experimental facility to minimize jitter.

An LED panel behind the camera provides a diffuse light source. The camera (Phantom® Miro 340,

Page 98: A Study on Heat Transfer in Dropwise Condensation

84

monochrome) is equipped with a macro lens and can capture images at up to 800 frames per second (FPS) at

1600 × 1600 resolution. In this imaging arrangement, pixels in the image correspond to 10 μm squares of the

test plate. Video data are used to qualitatively assess the condensation process and measure departing droplet

sizes (c.f. Section 4.2.3).

4.1.3 Sensor Plate and Amplification Module

Figure 4.5 Test plate: 1 – Gold-plated copper track used as an RTD in a 4-wire configuration; 2

– Gold finger contacts for connection to the RTD processing circuits; 3 – Area visible to the

camera (Figure 4.4-b).

Figure 4.5 presents a rendering of one side of the condenser test plate (Figure 4.1 (3)); the patterning

is identical on the back face. This RTD arrangement effectively provides space-averaged temperatures for 10

Page 99: A Study on Heat Transfer in Dropwise Condensation

85

zones on each side of the test boards. Temperature differences between corresponding sensor regions on both

faces of the boards can be used to determine average heat fluxes and HTCs for each zone.

Figure 4.6 a) Test plate dimensions in 𝐦𝐦. b) Stack-up of layers over the aluminum substrate

on the condensing side of the plate: (1) 100 nm thick conformal coat of parylene-C; (2) 80 nm

thick (nominal) ENIG® finish layer (3) 8 𝛍𝐦 thick copper tracks; (4) 85 𝛍𝐦 thick

COBRITHERM (Benmayor Aismalibar. proprietary) dielectric layer.

Page 100: A Study on Heat Transfer in Dropwise Condensation

86

Figure 4.6-a provides the dimensions of the test plate and Figure 4.6-b shows the stack-up of different

layers on the condensing side of the test plate. The stack-up is identical on the cooling side without the

conformal coat of hydrophobic parylene-C. The test plate is made out of 1.45 mm thick Aluminum 5052, which

is coated with an 85 μm nominal-thickness, high thermal conductivity dielectric on both sides (COBRITHERM

from Benmayor Aismalibar®, (4)). Measurements of the effective dielectric thermal conductivity are discussed

in section 4.1.4. The dielectric also serves as an adhesive that bonds the copper conductor layers to the aluminum

core. The copper tracks (3) that function as surface RTDs are nominally 8 μm thick and 127 μm wide. The net

resistance of each 4-wire RTD trace ranges from 25 Ω to 30 Ω. To prevent corrosion, the tracks are finished

with an Electroless Nickel Immersion Gold (ENIG®, (2)) process. A 100 nm thick layer of hydrophobic (θ =

94.8° [149]) Parylene-C (1) is applied on the condensing side to promote DWC using vapor deposition.

Page 101: A Study on Heat Transfer in Dropwise Condensation

87

Figure 4.7 a) Constant current supply and signal amplification board: 1 – Edge connector, which

mates with the test plate terminals; 2 – Signal amplification block; 3 – Constant current supply

block; 4 – Signal cable connector to the DAQ; b) Schematic of a single RTD, current and

amplifier loop: 1 – Constant current supply; 2 – Voltage signal amplification.

Figure 4.7-a shows a rendering of the constant current supply and signal amplification module, and

Figure 4.7-b shows a schematic of a single RTD, current supply, and voltage amplification loop. The gold

Page 102: A Study on Heat Transfer in Dropwise Condensation

88

fingers shown in Figure 4.5 mate directly into the edge connector. Two current and amplification boards are

installed, one on each side of the plate (10 RTD channels each). The current supply provides each RTD with

constant 5 mA current. This current is low enough that RTD self-heating (estimated at 300 W m−2) should be

less than 1 % of the steady state heat flux due to DWC. The amplifier scales the temperature dependent voltage

from each RTD by 40 × for processing. The amplified signal is sampled by the DAQ at 80,000 samples per

second (4,000 samples for each RTD per second). The current supply and amplifier circuits both employ passive

temperature compensation. The test plate, current supply, and DAQ are calibrated as a full integrated

temperature measurement system before tests.

4.1.4 Dielectric Thermal Conductivity Measurement

The test plate has thin layers of thermally conductive polymer dielectric on both faces to electrically

isolate the RTDs from the aluminum core. While the dielectric layers are relatively thin (85 μm, reported

𝑘D ~ 2 W m−1K−1), they still account for the major portion of overall thermal resistance through the test plate.

The manufacturer only reports a typical room temperature thermal conductivity of the dielectric. The high

temperature thermal conductivity of the dielectric and film thickness variation were unknown. To enable

accurate DWC heat fluxes and uncertainties, the following in-situ calibration study is performed to measure

effective thermal conductivity of the dielectric layers.

Page 103: A Study on Heat Transfer in Dropwise Condensation

89

Figure 4.8 Conduction resistance network schematic between surface temperature

measurements on the cold and the condensing faces of the test plate.

Figure 4.8 presents the thermal resistance network schematic between the surface temperature

measurements on the two sides of the test plate. Here, TC is the measured surface temperature on the cooled

face, RD,1 is the thermal resistance through the dielectric layer on the cooled side, RAl is the thermal resistance

through the aluminum core, RD,2 is the thermal resistance through the dielectric layer on the condensing side,

and Tsurf is the measured surface temperature on the condensing face.

A second test plate is used to measure the effective value of RD,2 (from the same production batch as

the main test plate). RTD temperature sensors are first calibrated per the procedure in the Appendix (A1). A

thermopile-based heat flux sensor (Fluxteq® PHFS − 01, 25 mm× 30 mm) is bonded to the condensing side

of the test plate with high thermal conductivity silicone adhesive, covering one of the central RTD elements.

This sensor is comprised of thin film metal sandwiched between layers of polyimide film. Equal thickness

polyimide film is applied around the periphery of the sensor to ensure nearly 1D conduction through the sensor.

Dropwise condensation experiments are then performed with this second plate. Local heat flux data from the

sensor are combined with surface RTD measurements to determine the thermal conductivity of the dielectric

layer kD through an inverse numerical method.

Page 104: A Study on Heat Transfer in Dropwise Condensation

90

Figure 4.9 COMSOL® Multiphysics mesh to determine thermal conductivity of the dielectric

layer: 1 – Coarse mesh; 2 – Fine mesh; 3 – Finest mesh.

COMSOL® Multiphysics [98] is used to perform the numerical portion of the calibration procedure. A

three-dimensional disc domain is used to simulate the aluminum substrate and the dielectric layers. The thin

copper traces are determined to introduce minimal thermal resistances, and are neglected. Figure 4.9 shows the

meshed domain, with different degrees of refinement. Steady state temperature readings from all the RTDs are

applied as fixed temperature boundary conditions over corresponding portions of the test plate faces. However,

the rectangular region covered by the heat flux sensor is instead modeled using a constant heat flux boundary

condition equal to the measured value.

The temperature dependent thermal conductivity of the aluminum substrate, kAl is provided as an input

to the numerical simulation. kD is then estimated by solving for the value that minimizes the absolute difference

between the measured RTD temperature under the flux meter and the average simulation value in the

corresponding region. The global optimal value of kD is found to be 0.937 W m−1 K−1, significantly lower than

the value reported by the manufacturer. Each enforced boundary condition is individually perturbed and the

corresponding sensitivities to kD are determined, yielding a propagated uncertainty accounting for RTD and

flux sensor errors of 0.046 W m−1K−1. Grid independence of the results is ascertained using the refinement

Page 105: A Study on Heat Transfer in Dropwise Condensation

91

process illustrated in Figure 4.9. It should be cautioned that this kD represents an effective value based on in-

situ calibration of the test plate at operating conditions and temperatures. The single value represents an average

for the cooled- and condensing-side dielectric layers, which are at different temperatures. This effective

conductivity may also incorporate effects due to variations in the dielectric layer thickness.

Page 106: A Study on Heat Transfer in Dropwise Condensation

92

4.2 Results and Discussion

4.2.1 Transient Heat Transfer Measurements

The experimental procedure begins with degassing of the steam loop using a vacuum pump and heating,

as described in section 4.1.1. The steam loop is sealed off to the outside environment after degassing. A port on

the test chamber is opened to the surroundings when the system pressure is greater than the ambient to facilitate

removal of non-condensables during tests. The test plate is electrically heated above the saturation temperature

and dried between tests. The sight glass is also heated to remove condensate for visual access.

Before each run, coolant water is pumped from the auxiliary tank (Figure 4.1 (6)), filtered, and

transferred to the main coolant tank (Figure 4.1 (5)). The main coolant tank is then sealed and pressurized with

compressed air (~ 400 kpa − gauge). Test plate temperatures, coolant temperature, steam generator

temperature and pressure, and steam side test section pressure and temperature are continuously monitored

during pre-experimental run preparations. The test plate is visually checked to confirm dryness. Each

condensation start-up experiment is not started until all the test plate is visibly clear of droplets and the plate

temperature is at least 5°C above Tsat.

Just before starting a condensation test, the test plate heater is turned off and test plate surface

temperature data logging is initiated. The solenoid valve (Figure 4.1) is opened to rapidly initiate spray cooling.

The high-speed camera (Figure 4.4) is triggered simultaneously with the solenoid valve to capture the DWC

startup.

Page 107: A Study on Heat Transfer in Dropwise Condensation

93

Figure 4.10 DWC on the vertical test plate at different times after startup (pressure = 1 atm)

Page 108: A Study on Heat Transfer in Dropwise Condensation

94

Figure 4.10 presents frames from high speed video of developing DWC, with cooling beginning at 𝑡 =

0 s. Images are of the portion of the test plate indicated in Figure 4.5. The condensing surface is vertical and

the system pressure is ~1 atm (103 kPa). At t = 0.50 s, the average wall temperature is Tsurf ~ 99.1 °C with

condensing-side heat flux q"surf ~ 250 kWm−2. The typical visible droplet radius is r ~ 150 μm. At t =

1.10 s, Tsurf ~ 97.2 °C, q"surf ~ 290 kWm−2, and r ~ 0.7 mm. Once droplets reach 𝑟max ~ 1.4 mm−

1.6 mm, they begin to slide down and depart the surface. In the t = 1.50 s and t = 1.90 s frames, it is apparent

how large sliding droplets “sweep” the surface of small droplets, allowing renucleation of small droplets in their

wake. These small droplets yield much higher heat fluxes per area than large droplets.

4.2.2 Heat Flux and Heat Transfer Coefficient Calculation Methodology

Figure 4.11 Schematic representing 1 dimensional discretization of the test plate, which is

divided into separate material zones with different material properties and discretization lengths:

1 – Dielectric zone (discretization length 𝚫𝐱𝐝); 2 – Aluminum zone (discretization length 𝚫𝐱𝐀𝐥); 3 – Dielectric zone (discretization length 𝚫𝐱𝐝); 4 – Boundary; 5 – Aluminum-dielectric interface;

6 – Aluminum-dielectric Interface; 7 – Boundary.

Page 109: A Study on Heat Transfer in Dropwise Condensation

95

An inverse approach is used to infer instantaneous surface heat flux and HTCs from the high frequency

surface temperature measurements. A finite volume solver is implemented to solve one-dimensional unsteady

heat transfer (Eqn. 4.1) for the test plate, comprising two dielectric layers sandwiching one aluminum layer

(Figure 4.11).

∂T

∂t=𝜕

𝜕𝑥(α∂T

∂x) 4.1

Here T is temperature, t is time, α is the temperature-dependent thermal diffusivity and 𝑥 is the spatial

coordinate through the plate thickness. The discretization is 2nd order accurate in space and 4th order accurate

in time (using the ODE45 integrator in MATLAB® [150]). The thermal conductivity k is harmonically

weighted at the interface using the discretization lengths on both sides of the interface (Eqn. 4.2).

kinterface =

Δxleft + Δxright

kleftΔxleft

+krightΔxright

4.2

Here left and right are with respect to the interface between the material zones in Figure 4.11.

Temperature readings are collected at 80,000 Hz for the RTDs (4,000 Hz for each RTD). The heat

transfer process is modeled as 1-D; therefore, measurements from two corresponding RTDs (one on the cooled

face and the other on the DWC face, covering the same areas) are used as transient temperature boundary

conditions for the finite volume solver. A steady state temperature profile based on the boundary temperature

Page 110: A Study on Heat Transfer in Dropwise Condensation

96

measurements from the RTDs is chosen as the initial condition. Temperature data are smoothed to remove noise

using a moving average filter (100-point window).

Heat flux, anywhere in the domain (Figure 4.11) can be calculated with

q" = k

Tn − Tn−1xn − xn−1

4.3

Here, Tn represents the temperature in a particular cell and xn represents its position.

HTCDWC is calculated using the following expression

HTCDWC = [TE − TSurfq"Surf

−thCkC]−1

4.4

It is important to note that the HTCDWCextracted from the experimental measurements includes the

effect of two additional physical resistances, namely: the promoter layer resistance and the constriction

resistance described in section 4.2.3.

Here q"surf is calculated using Eqn. 5.2 at the surface, TE (steam temperature) is derived from the

pressure readings described in section 4.2.1, Tsurf comes from direct measurements on the test plate surface,

thC is the parylene-C promoter layer thickness and kC is its thermal conductivity.

A grid independence study is performed to ensure sufficient resolution (Table 4.1).

Page 111: A Study on Heat Transfer in Dropwise Condensation

97

𝐍𝐃 𝐍𝐀𝐥 𝐀𝐛𝐬𝐨𝐥𝐮𝐭𝐞 𝐀𝐯𝐞𝐫𝐚𝐠𝐞 𝐃𝐞𝐯𝐢𝐚𝐭𝐢𝐨𝐧 % (𝐇𝐓𝐂𝐃𝐖𝐂) 5 15 -

10 30 0.5

20 60 0.1

Table 4.1 Grid independence calculation summary for the 1D, unsteady conduction equation

solver.

Here ND, in the first column, is the number of cells in the dielectric domain (Figure 4.11 (1,3)) and

NAl, in the second column, is the number of cells in the Aluminum substrate domain (Figure 4.11 (2)). Relative

deviation of HTCDWC, with respect to the previous grid resolution, is calculated at each time step. The Absolute

Average Deviation is then calculated using that data. Based on the summary given in Table 4.1, ND = 10 and

NAl = 30 are selected for the actual numerical calculations.

Figure 4.12 Temperature profile through the plate for time t = 0.3 s.

Page 112: A Study on Heat Transfer in Dropwise Condensation

98

Figure 4.12 presents the predicted temperature profile through the test plate at t = 0.3 s. The temperature

profile is non-symmetric, and therefore still evolving in time. This illustrates the need to use this inverse method

to extract instantaneous heat flux and HTC values instead of simpler steady state resistance formulae.

The test plate RTDs and the test chamber thermocouple (Figure 4.2 (9), which is used to calibrate the

pressure sensor – Figure 4.2 (8)) were calibrated using a high accuracy reference sensor (Accumac® AM8040,

±0.01 K). A conservative uncertainty is assumed for these instruments of ±0.2 K. A Monte Carlo method was

used to propagate these uncertainties along with the ~5% uncertainty in the dielectric thermal conductivity (kD,

c.f. Section 4.1.4) and ~5% uncertainty in the promoter layer thickness (thC) into the numerical model. 3 sets

of simulations (100, 200 and 400) with a random distribution of uncertainties (95% confidence) applied to the

temperature measurements, the dielectric thermal conductivity and the promoter layer thickness are run.

HTCDWC is calculated for each time step in each simulation. The bounds of 95% confidence are recorded (for

each time step) from the distribution of HTCs across the simulations. The uncertainty bounds are checked for

convergence. They converge sufficiently from 200 to 400 simulations. Uncertainty in HTCDWC, from the

condensation initiation (0.25 s) to 0.5 s, ranges from 20% to 8% and ranges from 8% to 5% after.

4.2.3 Validation Using Constriction Resistance and Theoretical Drop-Size Distribution Model

To validate the experimental system and approach, the steady state heat flux (q"surf) and the heat

transfer coefficient (HTC DWC) are compared with a theoretical drop-size distribution model, which accounts for

the constriction resistance due to the finite thermal conductivity of the condensation substrate and the thermal

resistance due to the presence of a promoter layer, which induces dropwise condensation on the surface.

Page 113: A Study on Heat Transfer in Dropwise Condensation

99

Constriction resistance in DWC was first discussed in detail by Mikic [151]. This theory argues that

the surface area under microscopic droplets experiences high heat transfer while that under larger droplets has

low flux except near the contact-line [36]. This uneven distribution of heat flux results in 3D conduction effects

in the substrate and constriction of the heat flow lines near areas populated with small droplets. The effective

conduction cross-section area is therefore smaller than the overall surface, analogous to contact resistance in

solids. An additional constriction resistance term can be defined in series with other transport stages to account

for this effect. Constriction resistance in DWC depends primarily on the thermal conductivity of the substrate

[63], [65], [152].

Figure 4.13 Schematic showing the constriction resistance 𝐑𝐬 and the promoter layer resistance

𝐑𝐂 between the surface temperature measured by the RTDs on the condensation side (𝐓𝐬𝐮𝐫𝐟) and

the Steam temperature (𝐓𝐄).

Figure 4.13 shows the schematic representation of thermal resistances (Constriction resistance RS,

Promoter layer resistance RC) between the temperature measured by the RTDs Tsurf on the test plate surface

(condensation side) and the steam temperature TE (TE = Tsat(PE)). These resistances have a considerable effect

on the overall heat transfer performance of a surface during DWC and the effect is discussed in this section.

These resistances are evaluated based on the theoretical models available in literature, which also serves to

demonstrate agreement between findings from this experiment and relevant prior work.

Page 114: A Study on Heat Transfer in Dropwise Condensation

100

The condensing substrate comprises of, from exposed face to bottom: a 100 nm thick conformal coat

of Parylene-C (DWC promoter), 100 nm thick ENIG® surface finish on the copper traces, 8 μm thick copper

traces, 85 μm thick dielectric layer, and 1.45 mm thick Aluminum 5052 substrate. Thermal resistance due to

the copper traces and their surface finish can be neglected as it is 2 orders of magnitude smaller than the smallest

other thermal resistance due to Parylene-C. To compare with the available data in literature, apparent thermal

conductivity of the condensing surface is calculated using the following expression

kA = (thAlkAl

+thdkd

+thCkC)−1

× (thAl + thD + thC) 4.5

Here thAl is the thickness of the aluminum substrate, thD is the thickness of dielectric layer on the

condensing side, and thC is the promoter layer thickness. kAl, kd, and kC are the thermal conductivities of

aluminum, dielectric and Parylene-C, respectively. The apparent thermal conductivity kA ~ 15 Wm−1K−1, is

comparable to that of stainless steels. Hannemann and Mikic [40] have reported HTCDWC ss = 62 kWm−2K−2.

Using HTCDWC Cu = 250 kWm−2K−2 [153] the value of additional constriction resistance for stainless steel

Rsss ~ 1.2 × 105 m2KW−1, which is used for calculations in the present study.

Beside constriction resistance, thermal resistance due to the promoter layer Rc, which induces DWC

on the surface, also has a significant impact on the heat transfer coefficient. To study the impact of Rc, we need

to look into heat transfer through a single droplet [36] in conjunction with the population balance theory [17],

[103], [133], which implies that at steady-state the number of droplets entering a size range must be equal to

the number of droplets leaving the size range.

Page 115: A Study on Heat Transfer in Dropwise Condensation

101

The steady-state heat flux through the condenser surface q"surftheory

can be calculated using the following

expression:

q"surftheory

= ∫ Qdrop(r)n(r)drre

rmin

+∫ Qdrop(r)N(r)drrmax

re

4.6

Here Qdrop(r) [W] is the heat transfer rate through a droplet of radius r, n(r) is the number of droplets

of radius r per unit area, per unit r, which grow directly due to condensation, N(r) is the number of droplets of

radius r per unit area, per unit r, which grow as a result of direct condensation as well as coalescence, and the

effective radius re is the transition radius between the two regimes of droplet growth. Heat transfer through a

single drop can be expressed as [154]:

Qdrop =

ΔTπr2(1 − rmin/r )

(1

2hi(1 − cosθ)+ F

θrk ∙ sinθ

+thC

kc sin2 θ)

4.7

Here ΔT = TE − Tsurf is the degree of surface sub-cooling, rmin is the minimum thermodynamically

viable radius for a droplet to grow, hi is the interfacial heat transfer coefficient, hi ~ 15.7 MWm−2K−1 at 1 atm

assumed here[24]. F(θ, r)[36] is a contact angle (θ) and droplet radius dependent function, which accounts for

the conduction heat transfer through a droplet. k is the thermal conductivity of the condensate, thC is the

thickness of the promoter layer, and kc is the thermal conductivity of the promoter layer.

The minimum drop radius, which depends on the degree of subcooling is given by [133], [153]

Page 116: A Study on Heat Transfer in Dropwise Condensation

102

rmin =

2σTsatρHlvΔT

4.8

Here σ is the surface tension of the condensate, Tsat is the saturated steam temperature, ρ is the density

of the condensate and Hlv is the liquid-vapor phase change enthalpy.

The effective radius re is defined as half the mean spacing between the active nucleation sites on the

condensing surface, assuming the nucleation sites form a square array. It is given by

re = (4Ns)

−1/2 4.9

where Ns is the number of active nucleation sites. Nucleation site density has been reported to be in

the range of 109 − 1015 and depends on the surface subcooling[133]. Rose[3] has proposed the following

expression for the site density

Ns =

0.037

rmin2 4.10

n(r) is the size distribution of small droplets growing as a result of direct condensation, and is calculated

using the population balance theory (Appendix, A2). N(r) for larger droplets was given by Le Fevre and Rose

[14]:

N(r) =

1

3πr2rmax(r

rmax)−2/3

4.11

Page 117: A Study on Heat Transfer in Dropwise Condensation

103

Here rmax is the experimentally determined departing droplet radius. Images from the high-speed

photography indicate rmax ~ 1.4 – 1.6 mm for the studied substrate and conditions.

Heat transfer coefficient HTCDWCtheory

can then be calculated as

HTCDWCtheory

=q"surftheory

ΔT 4.12

Here the minimum droplet radius rmin is given by Eqn. 4.8, which depends on the degree of

subcooling ΔT. For the same nucleation site density Ns ~ 2 × 1015 (Eqn. 4.10), δ = 100 nm, kc =

0.084 Wm−1K−1 (thermal conductivity of Parylene-C [149]), and adding an additional constriction resistance

Rsss, q"surf

theory and HTCDWC

theoryare calculated. Table 4.2 provides a comparison between the theoretically

calculated quantities and the mean steady state quantities from the experiment.

𝐪"𝐬𝐮𝐫𝐟𝐭𝐡𝐞𝐨𝐫𝐲

282 kWm−2

𝐪" 𝐬𝐮𝐫𝐟 268 ±10 kWm−2

𝐇𝐓𝐂𝐃𝐖𝐂𝐭𝐡𝐞𝐨𝐫𝐲

48 kWm−2K−1

𝐇𝐓𝐂 𝐃𝐖𝐂 46 ± 2 kWm−2K−1

Table 4.2 Comparison between theoretically calculated heat flux and

heat transfer coefficient with mean steady state quantities from the

experiment.

This provides a degree of validation for the experimental and post-processing approach developed in

this investigation.

Page 118: A Study on Heat Transfer in Dropwise Condensation

104

4.2.4 Discussion on Transient Heat Fluxes and Heat Transfer Coefficients

Figure 4.14 presents temperature readings from two corresponding RTDs on the two faces of the test

plate. The steam temperature TE is obtained from the pressure transducer in the test chamber (TE = Tsat(PE)).

At the high heat flux levels found in this representative test, quasi-steady temperature and heat flux are observed

after ~10 s (Figure 4.15-a). Quasi-steady readings after this startup period were used for comparison with

theoretical models in Section 4.2.3.

Transient data from this study has revealed three distinct phases of the startup period. In phase I, when

DWC initiates, the surface experiences high heat transfer due to the presence of small microscopic droplets.

The instantaneous HTC (Figure 4.16) reaches as high as 400 ± 50 kWm−2K−1. The HTC sharply declines

during this period as the mean droplet size grows.

In phase II, the droplets continue to grow, and the HTC declines, but at a more gradual rate than in

phase I. This transition might be attributed to a shift in droplet growth mechanism, i.e., from growth primarily

as a result of direct condensation to growth primarily due to coalescence. During this phase, droplets are still

small enough to be pinned by contact-line forces, and do not slide or depart. As the surface becomes covered

by relatively large droplets, the HTC falls to a minimum value of 27 ± 2 kWm−2K−1, which is even lower than

the mean steady state value of 46 ± 2 kWm−2K−1.

Phase III begins as the first droplets reach the critical departure size (1.4 – 1.6 mm) and begin sliding

downward. These droplets cause an avalanche effect, in which they collect other small to medium droplets in

their path. This clear the surface and initiate fresh nucleation of droplets on the surface. The first droplet

departures are somewhat synchronized, resulting in a spike in HTC up 90 ± 4 kWm−2K−1.

Page 119: A Study on Heat Transfer in Dropwise Condensation

105

Figure 4.14 a) Surface temperature on the Condensing side 𝐓𝐬𝐮𝐫𝐟, surface temperature on the

cooling side 𝐓𝐂 and the steam temperature 𝐓𝐄 during the experiment: 1 – First inflection point;

2 – Second inflection point; b) Magnified presentation of the chart area indicated above it.

Page 120: A Study on Heat Transfer in Dropwise Condensation

106

Figure 4.14-b presents a more detailed view of the surface temperature evolution during the early

stages of this DWC process (𝑡 = 0 − 3.5 s). This startup time is hypothesized to be a function of cooling

intensity. The present case corresponds to relatively high heat flux water spray cooling. An interesting question

to study in future would be the correlation between total cooling per unit area Qstartup, which is approximately

the amount of cooling (heat transfer [J m−2]) needed to condense the amount of liquid on the wall during steady

state, and the startup time. Integrating the heat flux on the cooling side q"C over the startup period for the present

case results in Qstartup ~ 1.1 MJ m−2. This is equivalent to 0.051 mL cm-2 of condensed liquid water (i.e., the

volume of water in a uniform film 510 μm thick).

The instantaneous heat flux (Figure 4.15), and the heat transfer coefficients (Figure 4.16), are inferred

using the computed unsteady temperature profiles through the substrate (Figure 4.12) at each time step. It is

important to note that the heat fluxes did not vary by more than 10% in different sections of the plate at any

time step, which confirms that the heat transfer is nearly 1 dimensional.

Figure 4.15 Solved heat fluxes on the cooled and condensing faces of the test plate.

Page 121: A Study on Heat Transfer in Dropwise Condensation

107

Figure 4.16 Evolution of heat transfer coefficients on the condensing face (a, 𝐇𝐓𝐂𝐃𝐖𝐂) and

cooled face (b, 𝐇𝐓𝐂𝐂). The insets show the uncertainty bounds for the HTCs.

Initially, the surface temperature is greater than TE due to the heating applied to dry the plate between

tests. Spray cooling is initiated on the back side of the plate at t = 0. The thermal conduction front reaches the

steam side of the plate after about 50 ms, which is comparable to the conduction time scale of the plate

(thAl2 αAl⁄ ) + 2(thD

2 αD⁄ ) = 58 ms. During the first ~ 0.25 s, the steam-side of the test plate is above the steam

saturation temperature, and no condensation is visible. The heat transfer coefficient on the cooling side varies

during this period as the spray flow stabilizes. The average HTCC = 1.5 ± .05 kW m−2 K−1 until t ~ 0.15 s.

HTC climbs to 27 ± 1 kW m−2 K−1 at t ~ 0.175 − 0.2 s, and begins to reduce toward the steady-state value

of 12 ± 0.5 kW m−2 K−1.

Page 122: A Study on Heat Transfer in Dropwise Condensation

108

At t ~ 0.25 s (Figure 4.14-b, point 1) Tsurf drops below Tsat and condensation begins. This can be

inferred from the decrease in the rate of surface temperature reduction. Three distinct phases are observed during

the startup. Based on the form of the surface temperature evolution (Figure 4.14-b), there is a period (0.25 –

1.15 s, points 1 → 2) of high condensation heat flux (phase I).In this phase HTCDWC goes from 360 ±

60 kWm−2K−1 to 60 ± 3 kWm−2K−1. It is followed by a period (1.2 – 2.5 s) of lower heat flux (phase II) in

which the HTC drops to 27 ± 2 kWm−2K−1. The HTC recovers (phase III) and reaches another local

maximum of 90 ± 4 kWm−2K−1 at ~ 3.2 s.

We hypothesize that the initially high heat flux is due to the surface being covered by microscopic

droplets, which yield higher heat transfer coefficients. The average droplet size increases, lowering heat transfer

from 1.2 – 2.5 s. Once some droplets become sufficiently large, their weight overcomes contact-line pinning

forces. Those droplets then slide down the surface, clearing small-to-medium size droplets and restarting

nucleation of microscopic droplets. The delay before droplets begin to slide, during which time the surface has

few droplets of the smallest size range, may explain the depression in heat flux between 1.2 – 2.5 s. This

inflection period highlights the need to characterize dropwise condensation start-up for thermal systems

operating in transient mode. Counter-intuitively, DWC heat transfer can be poor during part of the start-up

phase, because even though maximum droplets sizes are small, the departure and sweeping process has not yet

initiated.

An extension of our hypothesis implies that the instantaneous HTC could be inversely proportional to

the mean droplet radius (MDR) on the surface. Figure 4.17 shows MDR−1 plotted against the simulation time.

The data is from a numerical study analyzing growth of breath figures by Meakin [144]. The study uses Monte-

Carlo simulation technique to deposit fixed size droplets on a surface, at a random location one by one. If the

droplets touch each other, they merge. Droplets are merged till the surface has a droplet big enough to depart.

Page 123: A Study on Heat Transfer in Dropwise Condensation

109

The departing droplet slides down and captures any droplet in its path. No new droplets are added till the

departing droplet clears out of the surface. droplet deposition continues afterward, as described earlier. It is

important to note that the physics involving heat transfer and hydrodynamics of the droplets is not included in

the study. Nonetheless, it points at the three distinct phases of the startup process. The present work has captured

the first experimental verification of startup in DWC.

Figure 4.17 Inverse of the mean droplet radius S plotted against simulation time 𝐭∗. Adapted

from the numerical study by Meakin [144].

At the quasi-steady condition, the condensing side of the test plate averages about 6°C below TE

(average Tsurf ~ 94.4 °C and TE ~ 100.5 °C).

Page 124: A Study on Heat Transfer in Dropwise Condensation

110

4.3 Conclusions

Transient heat transfer process during dropwise condensation is studied experimentally. High temporal

resolution surface temperature and steam temperature measurements are carefully recorded. Unsteady heat flux

is calculated using a 1-D numerical inverse approach. Heat transfer coefficients are then determined, using

condensing side surface temperature and steam temperature. The main results from this study are as follows

• Transient heat transfer during dropwise condensation startup process, with high temporal resolution, is

studied for the first time. A uniquely instrumented test section and facility are designed to collect these

transient measurements.

• A high heat transfer spray cooling process is utilized in the present study and as a result, the startup period

lasts for only about 5 seconds. In contrast, another study by Parin et al.[41], which also reports HTC as a

function of time and utilizes a less effective cooling mechanism, doesn’t achieve steady state even after 700

s. Several studies [13], [24] in the past have reported the time to achieve steady state DWC on the order of

minutes.

• Three distinct phases of DWC startup process are identified and compared with the numerical data from

Meakin [144]. It is observed that the heat transfer can be poor during part of the startup process. Use of

DWC models for steady state operation would overestimate heat transfer during this period, leading to

undersized hardware designs.

Page 125: A Study on Heat Transfer in Dropwise Condensation

111

Chapter 5

Preliminary Study on Multi-Scale Simulation

of Dropwise Condensation

Page 126: A Study on Heat Transfer in Dropwise Condensation

112

One of the primary research challenges in studying droplet conduction is numerically predicting the

overall heat flux through the condensing surface. The active droplet size spectrum is wide, making direct

interface resolving simulations computationally unfeasible. For example, for water vapor condensing at Patm on

a surface cooled 5 K (Δ𝑇) below the saturation temperature, the smallest droplet radius 𝑟𝑚𝑖𝑛 =2𝜎𝑇𝑠𝑎𝑡

𝜌ℎ𝑙𝑣Δ𝑇=

2.59 × 10−9 m, and the largest droplet radius is ~O(10−3 m). There is a difference of 6 orders of magnitude

between the smallest and the largest drop, making direct simulation prohibitively expensive. Some investigators

have therefore developed fractal approaches to model the droplet distribution [112], [155], which allow

simulation of a smaller range of scales to predict the overall heat transfer in the larger domain. Others have used

a Lagrangian based droplet tracking framework [28] in which each droplet is treated like a rigid hemisphere

and individually tracked until it touches other droplets. Once the droplets come in contact they are

instantaneously replaced by a larger droplet (conserving the volume) in the next time step. This approach

requires tracking of only a few variables per droplet (e.g., position and radius). Lagrangian computing costs

may still be extreme for a practical-scale simulation as there are typically 107 – 109 active nucleation sites per

mm2 of the condensing surface [13]. One approach in Lagrangian studies has been to successively expand the

simulation area while reducing resolution in stages, as dropwise condensation evolves [28]. Heat transfer results

from finer stages can be applied to predict average heat transfer in the wall spaces between droplets in coarser

simulations. There can be multiple steps in between tracking the droplets on the smallest portion of the domain

to the entire domain. To date, none of these numerical studies of dropwise condensation have considered droplet

hydrodynamics during condensation. Such level of detail would require solution of fluid flow and heat transfer

equations with many individual degrees of freedom per resolved droplet. Interface tracking techniques like VOF

[121] or level set [156] can theoretically provide this level of detail, but due to computational cost, have only

been applied to study dynamics of few droplets [88], [90].

Page 127: A Study on Heat Transfer in Dropwise Condensation

113

The objective of the present study is to develop a modeling approach that captures the multi-scale

process of dropwise condensation to predict the overall transient heat transfer rate with acceptable

computational costs. In this approach, droplet dynamics for the larger (grid scale) droplets will be resolved with

an interface capturing approach (VOF). Heat transfer for the smaller (subgrid scale, SGS) droplets will be

predicted with an Eulerian averaged or Lagrangian SGS heat flux model (Figure 5.1). Smaller droplets are not

very mobile and do not directly affect droplets nearby on a large scale. Therefore, a relatively simple modeling

approach that captures SGS heat transfer, but neglects SGS hydrodynamics may be successful. Larger droplets

only directly contribute a small portion of the overall heat transfer (less than 10% of the total), but sweep clear

surface area during coalescence and sliding, effectively resetting the surface for fresh nucleation. Therefore, it

is important to capture their hydrodynamics.

Figure 5.1 Illustrative distribution of grid scale and subgrid scale drop sizes. Grid scale droplets

are resolved using VOF, and an average heat flux model is applied instead of resolving subgrid

scale droplets.

Page 128: A Study on Heat Transfer in Dropwise Condensation

114

Additional numerical and experimental data needs to be generated to develop and validate a robust SGS

average heat flux model. Preliminary simulation results for the approach are presented here. Challenges ahead

and additional work for the future are also discussed.

As described in Section 1.2 the current numerical approaches do not account for droplet history, and

the hydrodynamics involved in dropwise condensation. Techniques based on fractals are sensitive to specified

�� values (maximum droplet radius on the surface). They are quasi-steady in nature and therefore, not suitable

for predicting transient heat transfer, such as that during the DWC startup process. Techniques which are based

on droplet tracking have been shown to be insensitive to the initial droplet size distribution; Wu et al. [34] have

shown that their steady state drop size distribution and average heat flux predictions were the same whether the

initial droplets were of the same size or different. Predictions made by such models about the steady state drop

distribution may only be an emergent phenomenon, and they may not reliably predict transient behavior. The

proposed approach aims to combine empirical results from the experiments with direct simulation interface

tracking techniques to model dropwise condensation inexpensively, and accurately.

5.1 Modeling Approach

5.1.1 Grid-Scale Modeling

Grid scale droplets are resolved using VOF which is described in Section 3.2.1. However, in the

present study, the fluid generation term 𝜙 𝑝𝑐 which is given in Eqn. 3.1 is non-zero, as droplets are allowed to

Page 129: A Study on Heat Transfer in Dropwise Condensation

115

grow due to condensation. Eqn. 5.1 describes the dependence of the fluid generation term on the total heat

transfer through the surface.

��𝑝𝑐 =

(��𝑔𝑠 + ��𝑠𝑔𝑠)

𝜌ℎ𝑓𝑔

��𝑔𝑠 = ��𝑝𝑐

5.1

Here ��𝑔𝑠 is the grid scale volumetric heat transfer and ��𝑠𝑔𝑠 is the subgrid scale (SGS) volumetric heat

transfer (only applied in cells adjacent to the wall). The surface tension model, phase change model, and

dynamic contact angle model are the same as described in Section 3.2.

5.1.2 Subgrid Scale Modeling

Limited experimental transient heat flux data are available in dropwise condensation literature to guide

the modeling of subgrid scale average heat flux correctly. Experimental work in this dissertation will help

address this need in the future Chapter 4. In this preliminary proof of concept study, a representative SGS

model, adapted from the numerical study by Glicksman and Hunt [28], is used for the purpose of illustrating

the multi-scale simulation approach. Figure 5.2 shows the time dependent average heat flux data (𝑞"𝑠𝑔𝑠: curve

with the initial site density 108 cm−2) from [28] used to inform this SGS model.

Page 130: A Study on Heat Transfer in Dropwise Condensation

116

Figure 5.2 Average heat transfer coefficient vs time of a surface (given by Glicksman and Hunt

[28]).

The time dependent SGS model is applied to each cell neighboring the condensing surface, individually.

Each cell has a local clock for tracking of the evolving heat flux given by the model. The clock for each cell

which is swept over by a sliding grid scale droplet is reset to time = 0. ��𝑠𝑔𝑠 is also weighted by the phase fraction

term in each cell (Eqn. 5.2) so that it is only applied in the portion of each wall cell not covered by a grid-scale

droplet.

��𝑠𝑔𝑠 = ��𝑎𝑣𝑔 × 𝜙

��𝑎𝑣𝑔 = 𝑞"𝑠𝑔𝑠 × 𝐴𝑤𝑎𝑙𝑙 5.2

Page 131: A Study on Heat Transfer in Dropwise Condensation

117

Here 𝐴𝑤𝑎𝑙𝑙 is the area of the cell face neighboring the condensing surface boundary. The concept behind this

method is that, as the mesh is refined, the SGS spaces would be uncovered by grid-scale droplets for shorter

intervals of time. The SGS model would therefore account for a smaller range of the droplet size spectrum as

the mesh resolution is refined. If implemented consistently, this approach could yield mesh convergent

predictions for overall DWC heat transfer.

5.2 Illustrative Domain and Case Setup

A 10 cm × 10 cm× 1 cm domain is shown here as the test case (Figure 5.3). The test simulates a

condensing surface cooled 2 K below 𝑇𝑠𝑎𝑡 (373.15 K) at 𝑃𝑎𝑡𝑚 exposed to saturated steam. Droplet departure is

not considered. All boundaries are labelled in the figure, and the boundary conditions are given in Table 5.1.

Figure 5.3 Representative VOF domain for multi-scale simulation test case.

Page 132: A Study on Heat Transfer in Dropwise Condensation

118

Boundary Boundary condition

Symmetry

∇𝑇 ∙ �� = 0

∇𝑃 ∙ �� = 0

𝑈 ∙ �� = 0; ∇𝑈 ∙ �� = 0

∇𝜙 ∙ �� = 0

Atmosphere

∇𝑇 ∙ �� = 0

𝑃 = 𝑃𝑎𝑡𝑚

∇𝑈 ∙ �� = 0

∇𝜙 ∙ �� = 0

Wall (condensing surface)

𝑇 = (𝑇𝑠𝑎𝑡 − 1) K

∇𝑃 ∙ �� = 0

𝑈 = 0

�� ∙ ��𝑤 = cos(𝜃𝐷)

Table 5.1 Boundary conditions for the VOF domain.

Here �� is the normal vector (pointing in the liquid phase) to the interface, ��𝑤 is the normal vector

(pointing in the solid boundary) to the wall and 𝜃𝐷 is the dynamic contact angle, which is dependent on the 3-

phase contact line velocity.

A uniform initial temperature (𝑇𝑠𝑎𝑡) is applied to the entire domain. Here x and y are the directions

along the wall, and z is normal to the wall. Uniform grading 𝑔𝑟 =𝐿𝑛

𝐿1 is applied in the z direction. Here 𝐿1 is the

height of the layer of cells neighboring the condensing surface and 𝐿𝑛 is the height of the layer at the

“Atmosphere” boundary.

Page 133: A Study on Heat Transfer in Dropwise Condensation

119

5.3 Preliminary Results and Discussion

The simulation starts with a bare patch of wall. Liquid, in terms of phase fraction 𝜙, is initially

generated in each cell due to ��𝑠𝑔𝑠. As the simulation progresses the liquid accumulated in the cells is pulled

together to form tiny droplets due to the volumetric surface tension force implementation and the numerical

interface compression scheme[157]. These droplets grow due to condensation (Eqn. 5.1) and coalescence. Grid

scale droplets collect the liquid generated due to ��𝑠𝑔𝑠 on the bare surface between them. This is analogous to

tiny subgrid scale droplets coalescing with them. Figure 5.4-a shows the liquid coverage on the condensing

wall after about a minute of real time in the simulation. Several of the droplets in Figure 5.4-a are captured

during sliding and merging processes. Figure 5.4-b shows a 3D rendering of dropwise condensation on a

hydrophobic surface.

Figure 5.4 a) Liquid coverage on the condensing surface after the elapse of 1 minute in the

multi-scale simulation b) 3D rendering of another test simulation with contact angle set to

𝟏𝟎𝟎°.

Page 134: A Study on Heat Transfer in Dropwise Condensation

120

Figure 5.5 shows the cumulative surface coverage as percentage of the total condensing surface area

as a function of the droplet radius. The domain in this simulation is from a separate study and measures 0.5 cm

× 0.5 cm × 0.1 cm and the grid resolution is 500 × 500 in the XY plane. The drop distribution in the figure

evolves and approaches the distribution found in literature [14], [17].

Figure 5.5 Cumulative surface coverage by the grid scale droplets, of the condensing surface

vs droplet radii for different times.

Figure 5.6 shows the heat transfer maps from one of the many trial runs for the grid scale and the

subgrid scale separately. The grid scale heat transfer is mostly concentrated around the three-phase contact line

Page 135: A Study on Heat Transfer in Dropwise Condensation

121

of the droplets and therefore, the grid scale heat transfer map (Figure 5.6-a) traces the droplet outlines, as

expected. Since, the subgrid average heat flux model is time dependent, the subgrid scale heat transfer map

(Figure 5.6-b) traces droplet movement on the surface.

Figure 5.6 a) Grid scale heat transfer map; b) Subgrid scale heat transfer map.

This approach has to be refined and more rigorously assessed in conjunction with the experimental

results in Chapter 4. Robust, widely applicable VOF simulations for dropwise condensation can be set up with

the following steps:

• Apply the subcooling on the condensing surface.

• Apply the correct dynamic contact angle formulation, which depends on the surface-liquid pair, using

experimental data (c.f. section 3.1.4).

• Apply the transient heat flux from the experimental data (c.f. section 4.2.3) as an SGS model (section 5.1.2)

or use this to inform the selection of an SGS model. As the proposed SGS model form captures the evolution

Page 136: A Study on Heat Transfer in Dropwise Condensation

122

of small droplets on a portion of the surface after it is dried or cleared, transient startup heat flux data are

needed as a basis for this approach.

5.4 Conclusions

Earlier models for predicting the overall heat transfer through a surface treated the droplets as rigid

hemispheres or fractals, and did not resolve the fluid dynamics of the droplets. Due to the complex

hydrodynamics involved in dropwise condensation an interface capturing technique is needed to resolve the

droplets. Owing to the large spectrum of drop sizes, such a technique becomes quite expensive for resolving all

the scales.

A multi-scale approach is proposed here in which larger grid-scale droplets are resolved with the VOF

interface capturing technique and a subgrid scale average heat flux model is applied alongside to capture the

effect of smaller droplets. This approach needs a robust, widely applicable, and thoroughly validated subgrid

scale model. Experimental results described in this dissertation can serve as the primary subgrid scale heat

transfer model for such simulations. But some more work needs to be done in order to make these experimental

results generally applicable. Parametric studies, which characterize the effects of bulk substrate properties,

working fluid, surface properties and operating conditions need to be performed within the experimental

paradigm defined in chapter 4. For example:

• A parametric study to characterize the effect of cooling rate on DWC startup.

• A parametric study to catalogue dynamic contact angles for various solid-liquid pairs.

• A parametric study of startup at different pressure conditions etc.

Page 137: A Study on Heat Transfer in Dropwise Condensation

123

Chapter 6

Conclusions and Recommendations for

Future Research

Page 138: A Study on Heat Transfer in Dropwise Condensation

124

In this chapter major findings of this PhD dissertation on heat transfer in dropwise condensation are

summarized. Recommendations for future extension of this research are also suggested.

6.1 Numerical Simulations of dropwise condensation

Numerical techniques for predicting dropwise condensation heat transfer have long been sought.

Simulating the entire process using a fixed Eulerian grid is challenging as the drop size distribution spans 6

orders of magnitude and resolving all the length scales can get computationally expensive. Owing to these

challenges most of the studies in the past have either employed droplet tracking Lagrangian schemes simulating

a small (O(1 mm)) physical condensing surface area, or they have used fractal techniques, which assume the

drop size distribution to be similar at every length scale. Nearly all of these studies have assumed hemispherical

droplets with quasi-steady temperature profiles and have neglected the history of the droplets as they grow and

merge. Such approaches have generally used simplistic formulations for heat transfer through individual drops

and have treated the coalescence of drops as instantaneous. Chapter 2 of this dissertation presents the

development of an improved analytic model for predicting the quasi-steady heat transfer through individual

droplets valid for wide ranges of contact angles and relative interfacial and conduction resistances. Chapter 3

characterized the effect of droplet hydrodynamics during coalescence on condensation heat transfer. Specific

contributions and findings include:

• With the correct formulation of heat transfer through a drop, which takes into account the droplet size (Biot

number, Bi) and its contact angle (𝜃), steady state heat transfer predictions become more accurate. This

formulation can be used with a drop size distribution model to improve overall heat transfer predictions for

a condensing surface.

Page 139: A Study on Heat Transfer in Dropwise Condensation

125

• This formulation for individual droplet heat transfer can also be used to predict correct growth rates for

condensing droplets. It is shown in Chapter 2 that a prior widely used model for droplet heat flux resulted

in considerable error for droplet growth.

• Chapter 3 employed direct interface tracking simulations to explore the coalescence events during dropwise

condensation. This study illustrates how the complex hydrodynamics during coalescence of just 2 droplets

can have a significant effect on the overall heat transfer during dropwise condensation. Merging droplets

go through oscillations before they finally settle as 1 drop. During this period, they constantly sweep their

surrounding area and initiate fresh nucleation. Previous studies for modeling DWC have generally treated

coalescence as an instantaneous event without consideration oscillations or hydrodynamics. Findings here

suggest that such approaches significantly underpredict heat transfer during coalescence.

• Chapter 5 presents a multi-scale numerical framework for simulating dropwise condensation using a direct

interface tracking technique. Qualitative results are also presented to illustrate the grid-scale and the subgrid

scale heat transfer contributions during simulations. A path to reaching a complete framework is outlined.

6.2 Experimental Measurement of Heat Transfer During Dropwise Condensation

Startup

Prior to this investigation, there have been no reported experimental data for the transient process of

dropwise condensation startup at high heat fluxes relevant to industrial applications. Steady state descriptions

of DWC may be poorly suited for design of systems with cyclic or intermittent DWC processes, such as vapor

chambers for thermal devices with pulsed heat fluxes or refrigeration cycles operating for short periods of time.

Chapter 4 provides first experimental measurements during dropwise condensation startup.

Page 140: A Study on Heat Transfer in Dropwise Condensation

126

• Direct surface temperature measurements in conjunction with a finite volume discretization based, unsteady

conduction equation solver is used in this study to extract the transient heat transfer data (heat flux and

HTC) during DWC.

• Different aspects of the DWC startup process, like DWC initiation, droplet sliding initiation, and the quasi

steady state are observed and correlated with measured and calculated quantities, like surface temperature,

heat flux and HTC. Three distinct phases of the startup period are identified. In phase I, when DWC

initiates, the surface experiences high heat transfer due to the presence of small microscopic droplets ~O

(10−1) 𝜇𝑚 covering it. The drop in HTC is fast during this phase as the droplets grow rapidly. As the

droplets start to grow in phase II, the HTC goes down further. This drop in HTC is slower than phase I.

Sliding doesn’t begin in this phase and the droplets continue to grow, pinned to the surface, before reaching

their critical departure size. Phase III begins with the first sliding event quickly followed by the avalanche

of large droplets, which slide down colliding with other small to medium droplets in their path. They clear

the surface and initiate fresh nucleation of droplets on the surface. The above mentioned three distinct

phases characterize the startup in dropwise condensation and the inflection period in phase II emphasizes

the importance of this characterization for devices operating in transient mode.

• The results from the study can be used in guiding design decisions involving DWC applications. They can

also be utilized in formulating closure models for multi-scale VOF simulations.

6.3 Future Research Work

This dissertation provides important submodels, experimental data and methods, and a basic framework

to develop multi-scale simulation methods for dropwise condensation. Such simulations would directly resolve

Page 141: A Study on Heat Transfer in Dropwise Condensation

127

larger droplets using an interface tracking method and capture the effect of smaller droplets is captured using a

subgrid scale model. This approach should be advanced in the following ways:

• A robust, widely applicable, and thoroughly validated subgrid scale model is needed for DWC. Transient

experimental results described in this dissertation can serve to inform subgrid scale heat transfer models for

such simulations. More work is needed to expand the range of data on transient DWC startup. Parametric

studies, which characterize the effects of bulk substrate properties, working fluid, surface properties and

operating conditions need to be performed within the experimental paradigm defined in chapter 4.

• State of the art contact hysteresis models and surface tension force models can be integrated in this

framework to improve the fidelity of simulations. Implementation of these models is described in chapter

3. For example, the implementation of continuum surface tension force model developed by Brackbill et

al.[124] and contact angle hysteresis model described by Yokoi et al.[125].

• Chapter 4 hypothesizes the direct relation between the rate of cooling and the dropwise condensation startup

period. This prediction is based on observation. A parametric study to characterize the effect of cooling rate

will be a very useful design guideline for dropwise condensation applications.

Page 142: A Study on Heat Transfer in Dropwise Condensation

128

Appendix

A 1 Calibrating the Test Plate RTDs

Figure A.1 presents a schematic of the setup used to calibrate the RTDs on the test plate. The RTDs

are spread over a 75 mm diameter circular area, and a careful approach is needed to achieve nearly uniform

temperature over the whole plate during calibration to achieve the target ± 0.2 K uncertainty.

Figure A.1 Schematic cross-section representing the test plate RTD calibration setup: (1)

Reference temperature probe; (2) High thermal conductivity aluminum block; (3) intermediate

thermal conductivity stone block; (4) Temperature controlled heated plate; (5) – Thermal

insulation; (6) Signals to data acquisition system; (7) Test plate.

If the test plate were mounted directly on a temperature controlled hot plate, variations in heated plate

temperature (e.g., due to locations of individual heater elements) would be imposed on the test plate. To

minimize temperature variations in the test plate, a stack-up is made of a temperature controlled heated plate

(4), intermediate thermal conductivity stone (marble) plate (3), high conductivity thick aluminum plate (2),

Page 143: A Study on Heat Transfer in Dropwise Condensation

129

and the test plate (4).There’s thermal interface material (not shown in the figure) between each layer. The

whole system is insulated with thick mineral wool (5). In this arrangement, the stone plate acts a thermal

diffuser to spread out hot spots from the heated plate and the aluminum plate acts as a “ground plane” to

distribute heat over the stone plate, resulting in a nearly uniform temperature near the RTD and test plate. A

2D conduction finite element study was performed for this arrangement. The worst-case scenario of a linearly

increasing temperature profile (2 K range) is applied at the heated plate – stone plate boundary. The

temperature variation at the Aluminum block – test plate interface did not exceed 0.1 K.

This system was used to form piecewise linear calibration curves for the test plate RTDs with the

amplifier cards and data acquisition system at 5 temperatures from 30°C − 120°C, ensuring sufficient settling

time after each temperature adjustment.

Page 144: A Study on Heat Transfer in Dropwise Condensation

130

A 2 Population Balance Theory

Population balance implies that at steady state, the number of droplets entering a size range must be

equal to the number of droplets leaving the size range. For an arbitrary range of droplet radii, let r1 be the lower

bound and 𝑟2 be the upper bound (Δr = r2 − r1). Individual droplets grow at rate G = dr/dt. Population density

n(r) is defined as the number of droplets of base radius r per unit area per unit drop radius. The number of

droplets entering an arbitrary size range is A′n1G1dt and the number of droplets leaving it is A′n2G2dt. Here A′

is an arbitrary area on the condenser surface. Drops are also removed by the action of larger droplets sweeping

the area clean, and restarting the nucleation-growth process. The number of droplets swept off in a size range

is Sn12Δrdt, where S is the sweeping rate at which the condensing surface is renewed and n12 is the average

population density. To conserve the number of droplets in the size range

A′n1G1dt = A′n2G2dt + Sn12Δrdt 6.1

Eqn. 6.1 can be simplified as

A′(G2n2 − G1n1) = −Sn12Δr 6.2

As 𝛥𝑟 → 0, 𝑛12 → 𝑛. Therefore Eqn. 6.2 becomes

d(Gn)

dr+n

τ= 0 6.3

where sweeping period τ = A′/S .

Page 145: A Study on Heat Transfer in Dropwise Condensation

131

The heat transfer rate through a droplet (Qdrop) depends on interfacial resistance [122], [158],

conduction resistance [13], [28] through the drop, and the resistance of the promoter layer [60] under the droplet.

The combined effect of all the resistances on the heat transfer rate is given in Eqn. 4.7.

Heat transfer rate through a drop is also equal to the phase change enthalpy rate

Qdrop = ρHlv2πr

2(1 − cosθ)G 6.4

Using Eqns. 4.7 and 6.4 the droplet growth rate G can be written as a function of the droplet radius r

G = B1(1 − rmin/r)

B2r + B3

B1 =ΔT

2ρHlv

B2 =θ(1 − cosθ)

4k ∙ sinθ

B3 =1

2hi+δ(1 − cosθ)

kc sin2 θ

6.5

As G is a function of r, Eqn. 6.3 can be integrated with respect to r

′ ∫

d(Gn)

Gn= − ∫

dr

r

rmin

Gn

Gnmin

6.6

The minimum droplet radius rmin is given in Eqn. 4.8. Solving Eqn. 6.6 gives

Page 146: A Study on Heat Transfer in Dropwise Condensation

132

n(r) =(Gn)min

Gexp [−(

B2τB1

{(r2 − rmin

2 )

2+ rmin(r − rmin)

+ rmin2 ln(r − rmin)}

+B3τB1

{r − rmin + rmin ln(r − rmin)})]

6.7

The preceding expression of population density n(r) applies for small droplets that grow as a result of

direct condensation. For larger droplets, growing mainly due to coalescence, Le Fevre and Rose[14] gave an

expression for N(r) given in Eqn. 4.11. rmax is the radius of the largest droplet before it is removed by the body

forces. The latter is given by [2], [18]

rmax = K(σ

ρg)1/2

6.8

K is a parameter accounting for contact-angle hysteresis, which can be fit to experimental data (high-

speed photography data from the experimental run is used in its determination, K ~ 0.5 – 0.6), and g [ms−2] is

the gravitational acceleration.

The effective radius re is given by Eqn. 4.9.

In order to maintain continuity n(re) = N(re). Using this equality with Eqn. 6.7 and 4.9 gives (Gn)min.

Substituting (Gn)min into Eqn. 6.7 gives

n(r) =1

3πre3rmax

(rermax

)−23∙r(re − rmin)

(r − rmin)

(B2r + B3)

(B2re + B3)∙ exp[C1 + C2]

6.9

Page 147: A Study on Heat Transfer in Dropwise Condensation

133

C1 =B2τB1

{(re2 − r2)

2+ rmin(re − r) + rmin

2 log (re − rminr − rmin

)}

C2 =B3τB1

{re − r + rmin log (re − rminr − rmin

)}

At re d(log(n(r)))

d(log(r))=

d(log(N(r)))

d(log(r))= −

8

3 [60]. Using this expression, the sweeping period can now be

expressed as

τ =

3re2(B2re + B3)

2

B1(11B2re2 − 14B2rermin + 8B3re − 11B3rmin)

6.10

The steady state heat flux q"surftheory

, and the heat transfer coefficient HTCDWCtheory

can now be calculated

using Eqn. 4.6 and 4.12, respectively.

Page 148: A Study on Heat Transfer in Dropwise Condensation

134

Bibliography

[1] B. M. Burnside and H. A. Hadi, “Digital computer simulation of dropwise condensation from equilibrium

droplet to detectable size,” Int. J. Heat Mass Transf., vol. 42, no. 16, pp. 3137–3146, 1999.

[2] J. W. Rose and L. R. Glicksman, “Dropwise condensation-The distribution of drop sizes,” Int. J. Heat

Mass Transf., vol. 16, no. 2, pp. 411–425, 1973.

[3] J. W. Rose, “Further aspects of dropwise condensation theory,” Int. J. Heat Mass Transf., vol. 19, no. 12,

pp. 1363–1370, Dec. 1976.

[4] S. A. Stylianou and J. W. Rose, “Drop-to-filmwise condensation transition: Heat transfer measurements

for ethanediol,” Int. J. Heat Mass Transf., vol. 26, no. 5, pp. 747–760, 1983.

[5] S. A. Stylianou and J. W. Rose, “Dropwise Condensation on Surfaces Having Different Thermal-

Conductivities,” J. Heat Transf. Asme, vol. 102, no. 3, pp. 477–482, 1980.

[6] J. M. Beér, “High efficiency electric power generation: The environmental role,” Prog. Energy Combust.

Sci., vol. 33, no. 2, pp. 107–134, 2007.

[7] J. R. Lara, G. Noyes, and M. T. Holtzapple, “An investigation of high operating temperatures in

mechanical vapor-compression desalination,” Desalination, vol. 227, no. 1, pp. 217–232, 2008.

[8] T. B. Peters et al., “Design of an Integrated Loop Heat Pipe Air-Cooled Heat Exchanger for High

Performance Electronics,” IEEE Trans. Components, Packag. Manuf. Technol., vol. 2, no. 10, pp. 1637–

1648, Oct. 2012.

[9] J. J. Brunt and J. W. Minken, “The Application of Dropwise Promoters to Seawater Evaporators,” Ind.

Chem., vol. 34, no. 399, pp. 219–224, 1958.

[10] S. A. Abdul-Wahab, H. Al-Hinai, K. A. Al-Najar, and M. S. Al-Kalbani, “Fog Water Harvesting: Quality

of Fog Water Collected for Domestic and Agricultural Use,” Environ. Eng. Sci., vol. 24, no. 4, pp. 446–

456, May 2007.

[11] T. Nørgaard and M. Dacke, “Fog-basking behaviour and water collection efficiency in Namib Desert

Darkling beetles,” Front. Zool., vol. 7, no. 1, p. 23, 2010.

[12] K.-C. Park et al., “Condensation on slippery asymmetric bumps,” Nature, vol. 531, no. 7592, pp. 78–82,

Feb. 2016.

[13] J. W. Rose, “Dropwise condensation theory and experiment: A review,” Proc. Inst. Mech. Eng. Part A J.

Power Energy, vol. 216, no. 2, pp. 115–128, 2002.

[14] E. J. Le Fevre and J. W. Rose, “An Experimental Study of Heat Transfer By Dropwise Condensation,” Int.

J. Heat Mass Transf., vol. 8, no. 8, pp. 1117–1133, 1965.

[15] D. Beysens and C. M. Knobler, “Growth of Breath Figures,” Phys. Rev. Lett., vol. 57, no. 12, pp. 1433–

1436, 1986.

Page 149: A Study on Heat Transfer in Dropwise Condensation

135

[16] A. Umur and P. Griffith, “Mechanism of Dropwise Condensation,” J. Heat Transfer, vol. 87, no. 2, pp.

275–282, 1965.

[17] E. J. Le Fevre, “A Theory of Heat Transfer by Dropwise Condensation,” in Proceedings of the Third

Innternational Heat Transfer Conference, 1966, pp. 362–375.

[18] M. Abu-Orabi, “Modeling of heat transfer in dropwise condensation,” Int. J. Heat Mass Transf., vol. 41,

no. 1, pp. 81–87, 1998.

[19] B. T. Nijaguna, “Drop Nusselt Numbers in Dropwise Condensation,” Appl. Sci. Res., vol. 29, no. 1, pp.

226–236, 1974.

[20] J. L. Mccormick and E. Baer, “On the Mechanism of Heat Transfer in Dropwise Condensation,” J. Colloid

Sci., vol. 18, pp. 208–216, 1963.

[21] N. Fatica and D. L. V Katz, “Dropwise Condensation,” University of Michigan, 1949.

[22] J. W. Rose, “Dropwise condensation theory,” Int. J. Heat Mass Transf., vol. 24, no. 2, pp. 191–194, 1981.

[23] C. Graham and P. Griffith, “Drop size distributions and heat transfer in dropwise condensation,” Int. J.

Heat Mass Transf., vol. 16, no. 2, pp. 337–346, 1973.

[24] I. Tanasawa, “Dropwise Condensation - Progress Toward Practical Applications,” in Heat Transfer

Science and Technology, Proceedings of the International Symposium, 1987, pp. 43–53.

[25] C. Graham, “The Limiting Heat Transfer Mechanisms of Dropwise Condensation,” Masachusettes

Institute of Technology, 1969.

[26] I. Tanasawa and J. I. Ochiai, “Experimental Study on Dropwise Condensation,” Bull. JSME, vol. 16, no.

98, pp. 1184–1197, 1973.

[27] E. E. Gose, A. N. Mucciardi, and E. Baer, “Model For Dropwise Condensation On Randomly Distributed

Sites,” Int. J. Heat Mass Transf., vol. 10, pp. 15–22, 1967.

[28] L. R. Glicksman and A. W. Hunt, “Numerical Simulation of Dropwise Condensation,” Int. J. Heat Mass

Transf., vol. 15, no. 11, pp. 2251–2269, 1972.

[29] W. H. Wu and J. R. Maa, “On the Heat Transfer in Dropwise Condensation,” Chem. Eng. J., vol. 12, pp.

225–231, 1976.

[30] M. Mei, B. Yu, M. Zou, and L. Luo, “A numerical study on growth mechanism of dropwise

condensation,” Int. J. Heat Mass Transf., vol. 54, no. 9–10, pp. 2004–2013, 2011.

[31] F. Z. Sun et al., “The Fractal Dimension of the Fractal Model of Dropwise Condensation and its

Experimental Study,” Int. J. Nonlinear Sci. Numer. Simul., vol. 8, no. 2, pp. 211–222, 2007.

[32] C. Mu, J. Pang, Q. Lu, and T. Liu, “Effects of Surface Topography of Material on Nucleation Site Density

of Dropwise Condensation,” Chem. Eng. Sci., vol. 63, pp. 874–880, 2008.

[33] B. Qi, J. Wei, L. Zhang, and H. Xu, “A Fractal Dropwise Condensation Heat Transfer Model Including the

Page 150: A Study on Heat Transfer in Dropwise Condensation

136

Effects of Contact Angle and Dropsize Distribution,” Int. J. Heat Mass Transf., vol. 83, pp. 259–272,

2015.

[34] Y. T. Wu, C. X. Yang, and X. G. Yuan, “Drop distributions and numerical simulation of dropwise

condensation heat transfer,” Int. J. Heat Mass Transf., vol. 44, no. 23, pp. 4455–4464, 2001.

[35] S. Chavan et al., “Heat Transfer through a Condensate Droplet on Hydrophobic and Nanostructured

Superhydrophobic Surfaces,” Langmuir, vol. 32, no. 31, pp. 7774–7787, 2016.

[36] S. Adhikari, M. Nabil, and A. S. Rattner, “Condensation heat transfer in a sessile droplet at varying Biot

number and contact angle,” Int. J. Heat Mass Transf., vol. 115, pp. 926–931, 2017.

[37] H. Kim and Y. Nam, “Condensation behaviors and resulting heat transfer performance of nano-engineered

copper surfaces,” Int. J. Heat Mass Transf., vol. 93, pp. 286–292, 2016.

[38] R. Wilmshurst and J. W. Rose, “Dropwise Condensation: Further Heat Transfer Measurements,” in Fourth

International Heat Transfer Conference, 1970, pp. 1–11.

[39] H. Tanaka, “Measurements of Drop-Size Distributions During Transient Dropwise Condensation,” J. Heat

Transfer, vol. 97, no. 3, pp. 341–346, 1975.

[40] R. J. Hannemann and B. B. Mikic, “An experimental investigation into the effect of surface thermal

conductivity on the rate of heat transfer in dropwise condensation,” Int. J. Heat Mass Transf., vol. 19, no.

11, pp. 1309–1317, 1976.

[41] R. Parin et al., “Nano-structured aluminum surfaces for dropwise condensation,” Surf. Coatings Technol.,

vol. 348, pp. 1–12, 2018.

[42] E. Schmidt, W. Schurig, and W. Sellschopp, “Condensation of Water Vapour in Film and Drop Form,”

Tech. Mech. Thermodyn., vol. 1, pp. 53–63, 1930.

[43] W. A. Tanzola and J. G. Wiedman, “Film Forming Corrosion Inhibitors Also Aid Heat Transfer,” Pap.

Ind., vol. 36, pp. 48–50, 1954.

[44] P. Kajanne, “Experience with Filming Amines in Finland,” Corr. Prev. Control, vol. 4, pp. 49–53, 1957.

[45] A. Poll, C. J. Potter, and A. W. Powell, “Effect of Dropwise Condensation on Thermal Efficiency of a

Triple Effect Water Distillation Plant,” Proc. Instn. Mech. Engrs., vol. 181, no. 27, pp. 707–716, 1967.

[46] R. Wilmshurst and J. W. Rose, “Dropwise and Filmwise Condensation of Aniline, Ethanediol and

Nitrobenzene.,” Jpn Soc Mech Eng, p. p 269-273, 1974.

[47] Y. Utaka, A. Saito, H. Ishikawa, and H. Yanagida, “Study on Dropwise Condensation Curves (Dropwise

to Filmwise transition of Propylene Glycol, Ethylene Glycol and Glycerol Vapors on a Copper Surface

Using a Monolayer Type Promoter (Part I)),” Bull. Jap. Soc. Mech. Engrs., vol. 29, pp. 4228–4234, 1986.

[48] B. Misra and C. F. Bonilla, “Heat Transfer in the Condensation of Liquid Metal Vapor: Mercury and

Sodium up to Atmospheric Pressure,” in Chemical Engineering Prg. Symp. Ser., 1956, pp. 7–21.

[49] S. Necmi and J. W. Rose, “Heat-transfer measurements during dropwise condensation of mercury,” Int. J.

Page 151: A Study on Heat Transfer in Dropwise Condensation

137

Heat Mass Transf., vol. 20, no. 8, pp. 877–881, 1977.

[50] J. W. Rose, “Dropwise condensation of mercury,” Int. J. Heat Mass Transf., vol. 15, no. 7, pp. 483–488,

1972.

[51] J. Niknejad and J. W. Rose, “Dropwise Condensation of Mercury - Further Measurements,” in 6th

International Heat Transfer Conference, 1978, pp. 483–488.

[52] J. L. McCormick and J. W. Westwater, “Nucleation sites for dropwise condensation,” Chem. Eng. Sci.,

vol. 20, no. 12, pp. 1021–1036, Dec. 1965.

[53] J. L. Mccormick and J. W. Westwater, “Drop Dynamiccs and Heat Transfer During dropwise

Condensation for Water Vapor,” Chem. Engng. Prog. Symp. Ser, vol. 62, no. 64, pp. 120–134, 1966.

[54] A. C. Peterson and J. W. Westwater, “Dropwise Condensation of Ethylene Glycol,” Chem. Engng. Prog.

Symp. Ser, vol. 62, no. 64, pp. 135–142, 1966.

[55] D. W. Tanner, C. J. Potter, D. Pope, and D. West, “The effects of heat flux, steam velocity and non-

condensable gas concentration,” Int. J. Heat Mass Transf., vol. 8, pp. 419–426, 1965.

[56] R. Wilmshurst and J. W. Rose, “Dropwise and Filmwise Condensation of Aniline, Ethanediol and

Nitrobenzene.,” in Jpn Soc of Mech Eng, 1974, p. p 269-273.

[57] H. Tanaka and T. Tsuruta, “A Microscopic Study of Dropwise Condensation,” Int. J. Heat Mass Transf.,

vol. 27, no. 3, pp. 327–335, 1984.

[58] J. W. Rose, “Some aspects of condensation heat transfer theory,” Int. Commun. Heat Mass, vol. 15, pp.

449–473, 1988.

[59] J. W. Rose, I. Tanasawa, and Y. Utaka, “Dropwise Condensation,” in Handbook of Phase Change -

boiling and Condensation, Taylor and Francis, 1999, pp. 581–594.

[60] S. Kim and K. J. Kim, “Dropwise Condensation Modeling Suitable for Superhydrophobic Surfaces,” J.

Heat Transfer, vol. 133, no. 8, p. 081502, 2011.

[61] P. Griffith and M. S. Lee, “The effect of surface thermal propieties and finish on dropwise condensation.,”

Int. J. Heat Mass Transf., vol. 10, pp. 697–707, 1967.

[62] D. W. Tanner, C. J. Potter, D. Pope, and D. West, “Heat transfer in dropwise condensation - Part I The

effects of heat flux, steam velocity and non-condensable gas concentration,” Int. J. Heat Mass Transf., vol.

8, no. 3, pp. 419–426, 1965.

[63] D. G. Wilkins and L. A. Bromley, “Dropwise condensation phenomena,” AIChE J., vol. 19, no. 4, pp.

839–845, Jul. 1973.

[64] T. Takaharu, T. Hiroaki, and T. Shigenori, “Experimental verification of constriction resistance theory in

dropwise condensation heat transfer,” Int. J. Heat Mass Transf., vol. 34, no. 11, pp. 2787–2796, Nov.

1991.

[65] T. Takaharu and H. Tanaka, “A theoretical study on the constriction resistance in dropwise condensation,”

Page 152: A Study on Heat Transfer in Dropwise Condensation

138

Int. J. Heat Mass Transf., vol. 34, no. 11, pp. 2779–2786, 1991.

[66] Q. Zhao, D. Zhang, X. Zhu, D. Xu, Z. Lin, and J. Lin, “Industrial application of dropwise condensation,”

in Proceeding of the Ninth International Heat Transfer Conference, Hemisphere Washington, DC, 1990,

vol. 4, pp. 391–394.

[67] Q. Zhao, D. Zhang, X. Zhu, D. Xu, and J. Lin, “The Dropwise Condenser Design,” Nat. J., vol. 14, no. 3,

pp. 232–233, 1991.

[68] Q. Zhao et al., “Dropwise Condensation of Steam on Ion-Plating Surface,” in International Conference on

Petroleum Refining and Petrochemical Processing, 1991, pp. 1049–1052.

[69] Q. Zhao and J. Lin, “Research Progress on Industrial Application of Dropwise Condensation,” Chem. Ind.

Engng Prg., vol. 53, no. 2, pp. 17–20, 1991.

[70] Q. Zhao and D. Zhu, “Dropwise Condensation Surface with Corrosion Resistant Properties,” Corros. Rev.,

vol. 11, no. 1–2, pp. 97–103, 1993.

[71] Q. Zhao and W. Wang, “Vertical Dropwise Condensation Shell and Tube Heat Exchanger for Steam with

Water Cooling,” 91100529.3, 1993.

[72] Q. Zhao and B. M. Burnside, “Dropwise Condensation of Steam on Ion-implanted Condenser,” Heat

Recover. Syst. CHP, vol. 14, no. 5, pp. 525–534, 1994.

[73] Q. Zhao et al., “Dropwise Condensation of Steam on Vertical and Horizontal U-type Tube Condensers,” in

10th International Heat Transfer Conference, 1994, pp. 117–121.

[74] B. M. Burnside, M. P. Bonnar, H. A. Hadi, R. L. Reuben, J. I. B. Wilson, and Q. Zhao, “Dropwise

condensation in electricity utility turbine condensers - recent research work in China and U.K.,” in

Proceedings of the Universities Power Engineering Conference, 1996, vol. 2, pp. 673–675.

[75] G. O. Wang, Q. Zhao, G. Li, F. Fu, and S. Li, “Vertical Dropwise Condensation Shell and Tube Heat

Exchanger for Steam with Water Cooling,” 91201592.6, 1992.

[76] R. Fürstner, W. Barthlott, C. Neinhuis, and P. Walzel, “Wetting and self-cleaning properties of artificial

superhydrophobic surfaces,” Langmuir, vol. 21, no. 3, pp. 956–961, Feb. 2005.

[77] M. Miwa, A. Nakajima, A. Fujishima, K. Hashimoto, and T. Watanabe, “Effects of the surface roughness

on sliding angles of water droplets on superhydrophobic surfaces,” Langmuir, vol. 16, no. 13, pp. 5754–

5760, Jun. 2000.

[78] B. Qian and Z. Shen, “Fabrication of superhydrophobic surfaces by dislocation-selective chemical etching

on aluminum, copper, and zinc substrates,” Langmuir, vol. 21, no. 20, pp. 9007–9009, Sep. 2005.

[79] J. Y. Shiu, C. W. Kuo, P. Chen, and C. Y. Mou, “Fabrication of Tunable Superhydrophobic Surfaces by

Nanosphere Lithography,” Chem. Mater., vol. 16, no. 4, pp. 561–564, Feb. 2004.

[80] J. Genzer and K. Efimenko, “Recent developments in superhydrophobic surfaces and their relevance to

marine fouling: A review,” Biofouling, vol. 22, no. 5. pp. 339–360, Jan-2006.

Page 153: A Study on Heat Transfer in Dropwise Condensation

139

[81] L. Zhai et al., “Patterned superhydrophobic surfaces: Toward a synthetic mimic of the namib desert

beetle,” Nano Lett., vol. 6, no. 6, pp. 1213–1217, Jun. 2006.

[82] S. Wang, L. Feng, and L. Jiang, “One-step solution-immersion process for the fabrication of stable bionic

superhydrophobic surfaces,” Adv. Mater., vol. 18, no. 6, pp. 767–770, Mar. 2006.

[83] T. S. Wong et al., “Bioinspired self-repairing slippery surfaces with pressure-stable omniphobicity,”

Nature, vol. 477, no. 7365, pp. 443–447, 2011.

[84] N. Vogel, R. A. Belisle, B. Hatton, T. S. Wong, and J. Aizenberg, “Transparency and damage tolerance of

patternable omniphobic lubricated surfaces based on inverse colloidal monolayers,” Nat. Commun., vol. 4,

p. 2176, 2013.

[85] T.-S. Wong, T. Sun, L. Feng, and J. Aizenberg, “Interfacial materials with special wettability,” MRS Bull.,

vol. 38, no. 05, pp. 366–371, 2013.

[86] M. Mei, F. Hu, C. Han, and Y. Cheng, “Time-averaged droplet size distribution in steady-state dropwise

condensation,” Int. J. Heat Mass Transf., vol. 88, pp. 338–345, 2015.

[87] D. F. Cooper, B. H. Davey, and J. W. Smith, “Surface velocities during drop coalescence,” The Canadian

Journal of Chemical Engineering, vol. 54, no. 5. pp. 631–632, 1976.

[88] M. Sellier and E. Trelluyer, “Modeling the coalescence of sessile droplets,” Biomicrofluidics, vol. 3, no. 2,

p. 022412, 2009.

[89] Q. Liao, X. Zhu, S. M. Xing, and H. Wang, “Visualization study on coalescence between pair of water

drops on inclined surfaces,” Exp. Therm. Fluid Sci., vol. 32, no. 8, pp. 1647–1654, 2008.

[90] B. Singh Sikarwar, S. Khandekar, S. Agrawal, S. Kumar, and K. Muralidhar, “Heat Transfer Engineering

Dropwise Condensation Studies on Multiple Scales Dropwise Condensation Studies on Multiple Scales,”

Heat Transf. Eng., vol. 33, no. 33, pp. 4–5, 2012.

[91] J. Guadarrama-Cetina, R. D. Narhe, D. A. Beysens, and W. González-Viñas, “Droplet pattern and

condensation gradient around a humidity sink,” Phys. Rev. E - Stat. Nonlinear, Soft Matter Phys., vol. 89,

no. 1, 2014.

[92] C. M. Knobler and D. Beysens, “Growth Of Breath Figures on Fluid Surfaces,” Europhys. Lett., vol. 6, no.

8, pp. 707–712, 1988.

[93] A. Umur and P. Griffith, “Mechanism of Dropwise Condensation,” J. Heat Transf., vol. 87, no. 2, pp.

275–282, 1965.

[94] N. Miljkovic, R. Enright, and E. N. Wang, “Effect of droplet morphology on growth dynamics and heat

transfer during condensation on superhydrophobic nanostructured surfaces,” ACS Nano, vol. 6, no. 2, pp.

1776–1785, 2012.

[95] S. S. SadhalL and W. W. Martin, “Heat transfer through using differential drop condensate inequalities,”

Int. J. Heat Mass Transf., vol. 20, no. 12, pp. 1401–1407, 1977.

[96] N. Miljkovic, R. Enright, and E. N. Wang, “Growth dynamics during dropwise condensation on

Page 154: A Study on Heat Transfer in Dropwise Condensation

140

nanostructured superhydrophobic surfaces,” ASME 2012 3rd Int. Conf. Micro/Nanoscale Heat Mass

Transf. MNHMT 2012, pp. 427–436, 2012.

[97] S. Adhikari and A. S. Rattner, “Single droplet condensation heat transfer simulation data,” 2017. [Online].

Available: https://scholarsphere.psu.edu/concern/generic_works/z316q1768.

[98] “COMSOL, v. 5.2, Stockholm, Sweden.” Stockholm.

[99] K. C. Hayes and H. Hatze, “Passive visco-elastic properties of the structures spanning the human elbow

joint,” Eur. J. Appl. Physiol. Occup. Physiol., vol. 37, no. 4, pp. 265–274, 1977.

[100] I. B. Celik, U. Ghia, P. J. Roache, C. J. Freitas, H. Coleman, and P. E. Raad, “Procedure for Estimation

and Reporting of Uncertainty Due to Discretization in CFD Applications,” J. Fluids Eng., vol. 130, no. 7,

p. 78001, 2008.

[101] “MATLAB R2013, Natick, MA,” The Mathworks Inc. .

[102] M. Nabil and A. S. Rattner, “InterThermalPhaseChangeFoam-A framework for two-phase flow

simulations with thermally driven phase change,” SoftwareX, vol. 5, pp. 216–226, 2016.

[103] S. Vemuri and K. J. Kim, “An experimental and theoretical study on the concept of dropwise

condensation,” Int. J. Heat Mass Transf., vol. 49, no. 3–4, pp. 649–657, 2006.

[104] R. N. Leach, F. Stevens, S. C. Langford, and J. T. Dickinson, “Dropwise Condensation: Experiments and

Simulations of Nucleation and Growth of Water Drops in a Cooling System,” Langmuir, vol. 22, pp.

8864–8872, 2006.

[105] F. Chu, X. Wu, Y. Zhu, and Z. Yuan, “Relationship between condensed droplet coalescence and surface

wettability,” Int. J. Heat Mass Transf., vol. 111, pp. 836–841, 2017.

[106] J. E. Sprittles and Y. D. Shikhmurzaev, “Coalescence of liquid drops: Different models versus

experiment,” Phys. Fluids, vol. 24, no. 12, 2012.

[107] J. E. Sprittles and Y. D. Shikhmurzaev, “Finite element framework for describing dynamic wetting

phenomena,” Int. J. Numer. Methods Fluids, vol. 68, no. 10, pp. 1257–1298, 2012.

[108] R. Enright, N. Miljkovic, J. Sprittles, K. Nolan, R. Mitchell, and E. N. Wang, “How coalescing droplets

jump,” ACS Nano, vol. 8, no. 10, pp. 10352–10362, 2014.

[109] X. Chen, R. S. Patel, J. A. Weibel, and S. V. Garimella, “Coalescence-Induced Jumping of Multiple

Condensate Droplets on Hierarchical Superhydrophobic Surfaces,” Sci. Rep., vol. 6, no. November 2015,

pp. 1–11, 2016.

[110] B. J. Briscoe and K. P. Galvin, “Growth with coalescence during condensation,” Phys. Rev. A, vol. 43, no.

4, pp. 1906–1917, 1991.

[111] R. Narhe, D. Beysens, and V. S. Nikolayev, “Contact Line Dynamics in Drop Coalescence and

Spreading,” Langmuir, vol. 20, no. 4, pp. 1213–1221, 2004.

[112] M. Wu, T. Cubaud, and C. M. Ho, “Scaling law in liquid drop coalescence driven by surface tension,”

Page 155: A Study on Heat Transfer in Dropwise Condensation

141

Phys. Fluids, vol. 16, no. 7, pp. L51–L54, 2004.

[113] K. Rykaczewski, J. H. J. Scott, S. Rajauria, J. Chinn, A. M. Chinn, and W. Jones, “Three dimensional

aspects of droplet coalescence during dropwise condensation on superhydrophobic surfaces,” Soft Matter,

vol. 7, no. 19, pp. 8749–8752, 2011.

[114] Y. Nam, D. Seo, C. Lee, and S. Shin, “Droplet coalescence on water repellant surfaces,” Soft Matter, vol.

11, no. 1, pp. 154–160, 2015.

[115] S. T. Thoroddsen, K. Takehara, and T. G. Etoh, “The coalescence speed of a pendent and a sessile drop,”

J. Fluid Mech., vol. 527, pp. 85–114, 2005.

[116] Q. Liao, S.-M. Xing, and H. Wang, “Experimental investigation on coalescence of liquid drops on

homogeneous surface,” K. Cheng Je Wu Li Hsueh Pao/Journal Eng. Thermophys., vol. 27, no. 2, pp. 319–

321, 2006.

[117] F. Mashayek, N. Ashgriz, W. J. Minkowycz, and B. Shotorban, “Coalescence collision of liquid drops,”

Int. J. Heat Mass Transf., vol. 46, no. 1, pp. 77–89, 2003.

[118] G. D. Bansal, S. Khandekar, and K. Muralidhar, “Measurement of Heat Transfer During Drop-Wise

Condensation of Water on Polyethylene,” Nanoscale Microscale Thermophys. Eng., vol. 13, pp. 184–201,

2009.

[119] H. G. Weller, G. Tabor, H. Jasak, and C. Fureby, “A tensorial approach to computational continuum

mechanics using object-oriented techniques,” Comput. Phys., vol. 12, no. 6, pp. 620–631, 1998.

[120] A. S. Rattner and S. Garimella, “Simple mechanistically consistent formulation for Volume-of-Fluid based

computations of condensing flows,” J. Heat Transfer, vol. 136, no. 7, p. 071501, Mar. 2014.

[121] C. W. Hirt and B. D. Nichols, “Volume of Fluid (VOF) methods for the dynamics of free boundaries,” J.

Comput. Phys., vol. 39, no. 1, pp. 201–225, 1981.

[122] V. P. Carey, Liquid-Vapor Phase-Change Phenomena. Hemisphere, New York, NY (United States), 2008.

[123] D. Tam, V. von Arnim, G. H. McKinley, and A. E. Hosoi, “Marangoni convection in droplets on

superhydrophobic surfaces,” J. Fluid Mech., vol. 624, pp. 101–123, Apr. 2009.

[124] J. U. Brackbill, D. B. Kothe, and C. Zemach, “A continuum method for modeling surface tension,” J.

Comput. Phys., vol. 100, no. 2, pp. 335–354, 1992.

[125] K. Yokoi, D. Vadillo, J. Hinch, and I. Hutchings, “Numerical studies of the influence of the dynamic

contact angle on a droplet impacting on a dry surface,” Phys. Fluids, vol. 21, no. 7, 2009.

[126] M. Griebel and M. Klitz, “Simulation of droplet impact with dynamic contact angle boundary conditions,”

in Singular Phenomena and Scaling in Mathematical Models, Springer, Cham, 2014, pp. 297–325.

[127] T. Bergman, F. Incropera, D. DeWitt, and A. Lavine, Fundamentals of heat and mass transfer. 2011.

[128] S. Adhikari and A. S. Rattner, “Droplet Coalescence Heat Transfer Data,” 2018. [Online]. Available:

https://scholarsphere.psu.edu/concern/generic_works/bv73c0509.

Page 156: A Study on Heat Transfer in Dropwise Condensation

142

[129] N. Kapur and P. H. Gaskell, “Morphology and dynamics of droplet coalescence on a surface,” Phys. Rev.

E, vol. 75, no. 5, p. 056315, 2007.

[130] M. François, “Computations of Drop Dynamics With Heat Transfer,” University of Florida, 2002.

[131] C. J. Hurst and D. R. Olson, “Conduction through Droplets during Dropwise Condensation,” J. Heat

Transfer, vol. 95, pp. 12–19, 1973.

[132] J. L. Mccormick and E. Baer, Developments in Mechanics. Pergamon Press Ltd., 1965.

[133] J. W. Rose, “Dropwise condensation theory and experiment: A review,” Proc. Inst. Mech. Eng. Part A J.

Power Energy, vol. 216, no. 2, pp. 115–128, 2002.

[134] R. J. McGlen, R. Jachuck, and S. Lin, “Integrated thermal management techniques for high power

electronic devices,” in Applied Thermal Engineering, 2004, vol. 24, no. 8–9, pp. 1143–1156.

[135] J. M. Beér, “High efficiency electric power generation: The environmental role,” Prog. Energy Combust.

Sci., vol. 33, no. 2, pp. 107–134, 2007.

[136] Y. Chen et al., “Transparent superhydrophobic/superhydrophilic coatings for self-cleaning and anti-

fogging,” Appl. Phys. Lett., vol. 101, no. 3, p. 033701, Jul. 2012.

[137] S. G. Kandlikar, “A general correlation for saturated two-phase flow boiling heat transfer inside horizontal

and vertical tubes,” J. Heat Transfer, vol. 112, no. 1, pp. 219–228, 1990.

[138] N. Watanabe, M. Aritomi, and A. Machida, “Time-series characteristics and geometric structures of drop-

size distribution density in dropwise condensation,” Int. J. Heat Mass Transf., vol. 76, pp. 467–483, Sep.

2014.

[139] X. Liu and P. Cheng, “Dropwise condensation theory revisited Part II. Droplet nucleation density and

condensation heat flux,” Int. J. Heat Mass Transf., vol. 83, pp. 842–849, 2015.

[140] C. Gerardi, J. Buongiorno, L.-W. Hu, and T. Mckrell, “Measurement of nucleation site density, bubble

departure diameter and frequency in pool boiling of water using high-speed infrared and optical cameras,”

in ECI International Conference on Boiling Heat Transfer, 2009.

[141] N. Watanabe, M. Aritomi, and A. Machida, “Time-series characteristics and geometric structures of drop-

size distribution density in dropwise condensation,” Int. J. Heat Mass Transf., vol. 76, pp. 467–483, 2014.

[142] L. Redondo and J. F. Silva, “Solid state pulsed power electronics,” in Power Electronics Handbook,

Elsevier Inc., 2011, pp. 669–707.

[143] J. L. Hanson and L. L. Howland, “Method of operating a transport refrigeration unit,” US 5140826A, 11-

Jul-1992.

[144] P. Meakin, “Steady state behavior in a model for droplet growth, sliding and coalescence: the final stage of

dropwise condensation,” Phys. A Stat. Mech. its Appl., vol. 183, no. 4, pp. 422–438, 1992.

[145] Y. T. Wu, C. X. Yang, and X. G. Yuan, “Drop distributions and numerical simulation of dropwise

condensation heat transfer,” Int. J. Heat Mass Transf., vol. 44, no. 21–22, pp. 4455–4464, 2001.

Page 157: A Study on Heat Transfer in Dropwise Condensation

143

[146] S. Adhikari and A. S. Rattner, “Heat transfer during condensing droplet coalescence,” Int. J. Heat Mass

Transf., vol. 127, pp. 1159–1169, Dec. 2018.

[147] S. Ulrich, S. Stoll, and E. Pefferkorn, “Computer simulations of homogeneous deposition of liquid

droplets,” Langmuir, vol. 20, no. 5, pp. 1763–1771, 2004.

[148] D. W. Tanner, D. Pope, C. J. Potter, and D. West, “Heat transfer in dropwise condensation at low steam

pressures in the absence and presence of non-condensable gas,” Int. J. Heat Mass Transf., vol. 11, no. 2,

pp. 181–190, 1968.

[149] T. Goda, T. Konno, M. Takai, and K. Ishihara, “Photoinduced phospholipid polymer grafting on Parylene

film: Advanced lubrication and antibiofouling properties,” Colloids Surfaces B Biointerfaces, vol. 54, no.

1, pp. 67–73, Jan. 2007.

[150] “MATLAB, V. 9.6.0 (R2019a), Natick, Massachusettes.” The Mathworks Inc.

[151] B. B. Mikic, “On Mechanism of Dropwise Condensation,” Int. J. Heat Mass Transf., vol. 12, pp. 1311–

1323, 1968.

[152] T. Takaharu, T. Hiroaki, and T. Shigenori, “Experimental verification of constriction resistance theory in

dropwise condensation heat transfer,” Int. J. Heat Mass Transf., vol. 34, no. 11, pp. 2787–2796, 1991.

[153] I. Tanasawa, “Advances in condensation heat transfer,” Adv. Heat Transf., vol. 21, pp. 55–139, 1991.

[154] S. Kim and K. J. Kim, “Dropwise condensation modeling suitable for superhydrophobic surfaces,” J. Heat

Transfer, vol. 133, no. 8, Aug. 2011.

[155] B. M. Jose and T. Cubaud, “Droplet arrangement and coalescence in diverging/converging

microchannels,” Microfluid. Nanofluidics, vol. 12, no. 5, pp. 687–696, Mar. 2012.

[156] J. A. Sethian, Level Set Methods and Fast Marchoing Methods: Evolving Interfaces in Computational

Geometry, Fluid MEchanics, Computer Vision, and Materials Science, 3rd ed. Cambridge University

Press, 1999.

[157] H. G. Weller, “A new approach to VOF-based interface capturing methods for incompressible and

compressible flow,” 2008.

[158] S. S. Sadhal and M. S. Plesset, “Effect of Solid Properties and Contact Angle in Dropwise Condensation

and Evaporation,” in Transactions of the ASME, 1979, vol. 101, pp. 48–54.

Page 158: A Study on Heat Transfer in Dropwise Condensation

Vita

Sanjay Adhikari

Education

Ph.D., Mechanical Engineering, The Pennsylvania State University August 2020

M.S., Carnegie Mellon University December 2015

B.Tech., National Institute of Technology - Kurukshetra July 2011

Work Experience

Research Assistant, The Pennsylvania State University, University Park, PA Aug. 2015 – Aug. 2015

Graduate Intern, MAHLE Engine Components USA, Inc., Farmington Hills, MI Jan. 2015 – Jul. 2015

Systems Engineer, BrahMos Aerospace Pvt. Ltd., Hyderabad, India Jul. 2011 – Jul. 2013

Selected Publications

1. Adhikari, S, Nabil, M, Rattner, A S, “Condensation Heat Transfer in a Sessile Droplet at Varying Biot

Number and Contact Angle,” International Journal of Heat and Mass Transfer, 2017, Vol. 115, pp. 926-

931, DOI: 10.1016/j.ijheatmasstransfer.2017.07.077

2. Adhikari, S, Rattner, A S, “Heat Transfer During Condensing Droplet Coalescence,” International

journal of Heat and Mass Transfer, 2018, Vol. 127, pp. 1159-1169, DOI:

10.1016/j.ijheatmasstransfer.2018.07.005


Recommended