You are on page 1of 9

Home

Search

Collections

Journals

About

Contact us

My IOPscience

Droplet ejection performance of a monolithic thermal inkjet print head

This article has been downloaded from IOPscience. Please scroll down to see the full text article.
2007 J. Micromech. Microeng. 17 1420
(http://iopscience.iop.org/0960-1317/17/8/002)
View the table of contents for this issue, or go to the journal homepage for more

Download details:
IP Address: 131.114.31.96
The article was downloaded on 30/11/2011 at 10:46

Please note that terms and conditions apply.

IOP PUBLISHING

JOURNAL OF MICROMECHANICS AND MICROENGINEERING

doi:10.1088/0960-1317/17/8/002

J. Micromech. Microeng. 17 (2007) 14201427

Droplet ejection performance of a


monolithic thermal inkjet print head
A K Sen and J Darabi
MEMS and Microsystems Laboratory, Department of Mechanical Engineering,
University of South Carolina, 300 Main Street, Columbia, SC 29208, USA
E-mail: darabi@engr.sc.edu

Received 2 January 2007, in final form 20 May 2007


Published 21 June 2007
Online at stacks.iop.org/JMM/17/1420
Abstract
This paper presents a simulation study of the droplet ejection performance
of a thermal inkjet print head. The geometry of the print head comprises a
dome-shaped ink chamber, a nozzle guide and a ring-shaped heater
integrated on each chamber. The design eliminates direct contact between
the heater and the ink, thus minimizing heater burnout. The ink manifold,
ink chamber and nozzle are aligned, thus facilitating higher nozzle density.
The model simulates thermal bubble dynamics including nucleation and
growth of thermal bubbles caused by a thermal pulse. The model was
validated by comparing model predictions with experimental results for a
previously reported print head design. Then, the model was used to simulate
the droplet ejection performance of the proposed inkjet print head. Effects
of print head geometry including nozzle diameter, nozzle length, chamber
size, heater dimensions and location, thermal conductivity of the passivation
layer, operating conditions including total thermal energy and pulse width,
properties of the ink including density, viscosity and surface tension on the
performance of the inkjet device are investigated. The influence of these
parameters on the drop volume and velocity, threshold energy and tail length
of the ejected droplets is studied.
(Some figures in this article are in colour only in the electronic version)

1. Introduction
Inkjet printing is considered one of the most promising printing
technologies that offers several advantages including high
speed, quiet operation and compatibility with a variety of
substrates [1]. The technology has become increasingly
important because of its compatibility with low-cost computer
printers [2]. The initial printers were expensive and had
reliability issues due to difficulties associated with clogging.
A significant improvement in the technology occurred in 1985
with the development of the Thinkjet printer by HewlettPackard, which could be batch fabricated, offering low-cost
and reduced reliability issues [3]. Following this, inkjet
printers have improved continually and established a large
market share due to their low cost, high resolution and low
noise [4]. A variety of inkjet printing mechanisms have
been developed including thermal inkjet (TIJ), piezoelectric
and electrostatic. TIJ has been the preferred method of
inkjet printing due to its low cost and high quality [5].
0960-1317/07/081420+08$30.00

TIJ technology has a greater market share for printers than


any other digital printing technology and all other inkjet
technologies combined. A TIJ printing system utilizes a heater
element and heat pulse to locally heat the ink present in a
chamber to generate vapor bubbles. These bubbles grow with
time, resulting in ejection of the ink droplets from the chamber
through a nozzle. When the ink droplet is ejected, the heat
pulse is removed causing the bubble to collapse due to which
ink is retracted into the chamber. The most important part of a
TIJ printing system is the print head which comprises the ink
chamber, inlet manifold, a nozzle and heaters.
Several researchers have reported the development of
thermal inkjet print head designs. Lee et al [3] proposed
and demonstrated a monolithic process to fabricate a dropon-demand type TIJ print head. Tseng et al [6] introduced
an innovative microinjector design employing a bubble valve,
which results in superior droplet ejection characteristics. They
described fabrication, implementation and characterization of
a thermal driven micro-injector. Lee et al [7] described

2007 IOP Publishing Ltd Printed in the UK

1420

Droplet ejection performance of a monolithic thermal inkjet print head

a thermally-driven inkjet print head that consists of domeshaped ink chambers and thin film nozzle guides. Tseng et al
[8] described a thermally driven droplet injector employing
a novel concept of virtual chamber neck. Wang et al
[9] developed a monolithic thermal bubble inkjet print head
based on silicon micromachining technology. Park and Oh
[10] presented a thermally driven monolithic inkjet print
head comprising dome-shaped ink chambers, thin film nozzle
guides and heaters integrated on the top surface of each
chamber. Bae et al [11] proposed a TIJ printer head on a
SOI wafer that comprised two rectangular heaters. Baek et al
[12] proposed a novel high-density TIJ to overcome defects
due to the limitations of the nozzle density in conventional
TIJs. Huang et al [13] designed and fabricated a monolithic
print head that utilizes back-shooting bubble nucleation for
ejection of droplets. Development of a low cost, high
resolution and reliable inkjet print head is critical for the
future of inkjet printing. A high resolution can be achieved
by reducing the nozzle pitch and diameter, hence the droplet
size. Novel design concepts can improve the print quality by
eliminating formation of satellite droplets. The inkjet print
heads can be made cost effective by reducing the overall size
of the print head but the trend goes to larger print heads
for achieving higher printing speed. Most of the existing
inkjet print head designs utilize assembly or bonding processes
for integrating nozzle plates with ink chamber. Simpler
design and fabrication approaches should be adopted to reduce
cost. Monolithic inkjet print heads have been developed with
significant advantages over conventional print heads in terms
of fabrication cost, resolution and reliability [7].
The physics of the TIJ printing process involves many
physical phenomena including heat conduction, bubble
nucleation, two-phase flow, fluid dynamics and surface
chemistry. The most important performance parameters
include droplet speed, droplet volume, operating frequency
and thermal energy. Development of new inkjet devices or
improvements in existing designs require a multiple cycle
of redesign, fabrication and testing to achieve design goals.
As an economical alternative, the performance of inkjet print
heads can be accurately predicted using numerical models.
Numerical simulation can be used to study the effects of design
parameters including heater layer structure, chamber geometry
and operating conditions. A number of studies have been
reported on numerical modeling of ink ejection process in TIJs.
Asai et al [14] reported a one-dimensional linear hydraulic
model to predict bubble-growth and liquid motion coupled
with heat transfer in a side-shooter head. Chen et al [15]
modeled the ink ejection process in a top shooter head using a
lumped model. In this paper, we present the performance of a
monolithic thermal inkjet print head by performing numerical
simulation using FLOW3D. The geometry of the print head
comprises a dome-shaped ink-chamber, a nozzle guide and a
ring-shaped heater integrated on the surface of each chamber.
The print head design eliminates direct contact between the
heater and the ink, thus minimizing heater burnout. The
ink manifold, ink chamber and the nozzle are aligned, thus
facilitating higher nozzle density. A similar inkjet print head
design based on an omega-shaped heater has been reported by
Park and Oh [10], but the goal was to investigate the durability
of the inkjet print head through experiments.

2. Theory
In a TIJ print head, the droplet ejection mechanism involves
coupled interaction between the fluid flow and heat transfer.
The model considers heat transfer from the heater to the
ink present in the ink chamber through a passivation layer.
The temperature dependence of the fluid properties including
viscosity, density and surface tension is taken into account.
Bubble dynamics theory is used to simulate the nucleation
and growth of the thermal bubbles. The flow of ink exiting
the nozzle and its further motion after exiting the nozzle is
simulated. The free surface between the ink and the medium
(air) is tracked using volume of fluid method. The fluid
dynamic behavior of drop ejection in an incompressible,
Newtonian fluid with viscosity , density , is governed by
the continuity and the NS equations with specified boundary
conditions:
u
"=0
d"
u
= p + 2 u
" + g

dt

(1)
(2)

where u is the flow velocity, p is the fluid pressure, g is gravity


and d/dt is the Lagrangian derivative. The balance of forces
along the fluidair interface is governed by the mechanical
boundary condition
(3)
n" [T" m ] n" = n"

where Tm is the mechanical stress tensor, is the surface


tension of the fluid and n is the unit normal of the interface
directed into the fluid medium. The heat transfer in the solid
and in the fluid is solved using energy equation,
DT
(4)
= k 2 T
cp
Dt
where T is the temperature, D/Dt is the material derivative,
Cp is the specific heat and k is the thermal conductivity. The
nucleation and growth of thermal bubbles is simulated using
bubble dynamics theory. The equation-of-state for bubbles is
represented by,
pb = ( 1)g cvvap Tv

(5)
vap

where pb is the bubble pressure, g is the gas density, Cv


is the specific heat at constant volume, and Tv is the vapor
temperature in absolute scale, is the specific heat ratio for
the gas. As the bubble formation occurs through phase change,
the saturation pressure of the vapor Psat is expressed in terms of
its temperature T through the Clapeyron equation as follows,
! !
"#
vap "
1
( 1)cv
1
P sat = pv1 exp

(6)
T
tv1
$H
where ( pv1, tv1) is a point on the saturation curve, and $H
is latent heat of evaporation. The rate of phase change is
proportional to the deviation from saturation conditions. The
formulation to compute the rate of phase change based on
kinetic theory is [16]
$
%
&
M
P sat
Pv
cevap l ccon
(7)
m=
2R
Tv
Tl
where m is the net mass transfer, M is the molecular
weight of vapor, R is the vapor gas constant, Tl and Tv
are the temperatures in the liquid and vapor. Plsat indicates
1421

A K Sen and J Darabi

Heater

Passivation layer

t
D

Ink chamber
Nozzle guide

Ink channel

L
Figure 2. Computational model of the inkjet print head.

Ink manifold

Figure 1. A cross-sectional view of the inkjet print head.

saturation pressure corresponding to liquid temperature Tl,


coefficients cevap and ccon are the accommodation coefficients
for evaporation and condensation, respectively. The energy
changes in the fluid are computed as follows. If $m is the
mass exchange between fluid and vapor in an element, the
corresponding change in the liquid energy in the element is
'
'$m(H + e ),
if $m ! 0
lv
l
'
(
(8)
$E = '
vap )
'$m Hlv + cv Tv ,
if $m " 0

where Hlv denotes the latent heat of vaporization and el is the


specific energy of the liquid at temperature Tl. The fluidair
interface is tracked using the volume of fluid (VOF) technique
[17]. The motion of the interface between the liquid and air is
defined by a volume fraction function F in equation (9), and
the interface is tracked using the conditions in equation (10):
dF
+u
" F = 0
dt

outside the liquid


0
inside the liquid
F (x, y, z, t) = 1

>0, <1 on the free surface.

(9)
(10)

Equation (9) also represents the kinematic condition at the


interface. The simulation tool solves equations (1)(10) with
appropriate boundary conditions, as applicable to any specific
print head design.

3. Simulation of the proposed TIJ print head


3.1. Description of the device structure
A cross-sectional view of the structure of the proposed TIJ
print head is depicted in figure 1. The device comprises an
ink manifold, dome-shaped ink chamber, a thin film nozzle
guide, ring-shaped heaters and a passivation layer in between
the heater and the ink chamber. The design has the ink
chamber and the nozzle inline, thus higher nozzle density can
be achieved. It includes a passivation layer between the heater
and the ink chamber thereby eliminating heater burn out issues.
The shape of the ink chamber is hemispherical (dome-shaped)
which can be fabricated on a Si-substrate using isotropic wet
bulk micromachining. The nozzle guide and the heater can be
formed using thin film deposition.
1422

3
Figure 3. Boundaries of the computational domain.

3.2. Description of the computational model


A schematic diagram of the computational model of the inkjet
print head simulated in the present study is depicted in figure 2.
The diameter of the ink chamber is D, the nozzle diameter is d,
the heater width is w and the thickness of the passivation layer
is t. x defines the location of the heater w.r.t. the nozzle. The
width of the thermal pulse applied to the heater element is .
The printing surface is located at a distance L from the nozzle
orifice. The ink present inside the ink chamber is of density ,
viscosity and surface tension . The thermal conductivity
of the passivation layer is k.
3.3. Boundary conditions
A two-dimensional axi-symmetric model is considered for
simulation. The boundary conditions for the model are
illustrated with reference to figure 3. Boundary 1 (inlet
boundary)stagnation pressure boundary conditions that
represent the external ambient conditions as a large reservoir
of stationary fluid, and fluid fraction function F = 1, boundary
2symmetry boundary condition which indicates that the
same physical processes exit on the two sides of the boundary,
boundary 3 and 4continuative boundary condition which
represents zero normal derivative at the boundary for all
quantities, F = 0, all the solid boundaries: no-slip boundary
conditions. The contact angle used in the simulation was
10 which represents the full-wetting condition. The initial
temperature in the computational domain is 300 K.
3.4. Model validation
The numerical model is validated by comparing the model
predictions with the experimental data and results of a lumped
model for the axi-symmetric TIJ print head reported by Lee
et al [18]. The droplet momentum as a function of nozzle
diameter predicted by the present model is compared with
that measured in the experiments and obtained from the
lumped model in figure 4. The agreement between the model
predictions and the experimental data is within 10% indicating
the validity of the model and modeling methodology.

Droplet ejection performance of a monolithic thermal inkjet print head

12
Droplet speed

Lumped model [19]

200

175

150
15

20
25
Nozzle length (m)

30

Figure 4. Comparison of model predictions with the experimental


data and results of the lumped model [18].

Figure 5. A 2D axi-symmetric mesh with 90 radial and 300 axial


cells in the computational domain.

3.5. Results and discussion


The geometry of the model (figure 2) considered in the
simulation has the following dimensions: D = 50 m, d =
18 m, w = 10 m, t = 2 m. The printing surface
is located at a distance of 0.5 mm from the nozzle orifice.
The passivation layer is SiO2 with a thermal conductivity of
1.5 W mK1, specific heat 740 J kg1 K1 and density
2200 kg m3. The properties of the ink used in the simulation
are as follows: density 1500 kg m3, viscosity 0.001 Pa s
and surface tension 0.02 N m1. A thermal pulse of 3 s
period (pulse width) and 4 W power is applied to the heater
surface. The model is axi-symmetric for which a 2D axisymmetric mesh is built around the model with 90 radial and
300 axial cells as depicted in figure 5. To ensure that the results
reported herein are mesh independent, a mesh convergence
study was performed. The droplet volume and speed (at a
distance of 0.5 mm from the nozzle orifice) as functions of
the total number of cells in the computational domain are
studied and the results are presented in figure 6. The results
show that mesh convergence is achieved for a total number
of cells higher than twenty-thousand in the computational
domain. Time evolution of the fluid interface together with
the temperature contours in the fluid is depicted in figure 7.
A thermal pulse applied to the heater generates a rapid increase
in the heater temperature. The heat is transferred to the
fluid in the chamber through the passivation layer. The fluid
film next to the passivation layer and near the rapidly-heated

9
8

6
7
3

0
5000

Droplet speed (m/s) a

Expe riment al [19]

225

10
Droplet volume

Model predictions

Droplet volume (pl) a

Droplet momentum (ng.m/s)a

250

10000 15000 20000 25000


Total number of cells

5
30000

Figure 6. Droplet volume and speed as functions of total number of


cells in the computational domain.

heater element is subjected to an extremely high temperature


sufficient to initiate bubble nucleation. The bubbles expand
subsequently with time pressurizing and displacing the fluid
from the chamber. The growing bubbles begin to collapse
when heat transfer to the fluid is reduced due to decrease in
the heater temperature.
The variation of the maximum temperature of the heater
and the fluid film (fluid film enclosing the bubble) with time is
presented in figure 8. The temperatures of the heater and the
fluid film reach their maximum values and then decrease with
time. The temperature of the heater attains its maximum value
at the end of the thermal pulse (3 s). However, the maximum
temperature of the fluid film occurs at a later time step (5 s)
due to the time required for heat transfer to take place through
the passivation layer. At some instant the bubbles begin
to collapse and the fluid in the chamber is retracted toward
the collapsing bubbles. A fluid meniscus is formed inside
the nozzle which provides the capillary force necessary for
refilling the chamber.
The fluid is ejected through the nozzle orifice in the form
of a primary droplet dragging a long tail. The tail is then pulled
back into the primary droplet due to surface tension resulting
in a single droplet. Interestingly, no satellite droplets are
observed. Thus the design eliminates the formation of satellite
droplets, a frequent problem in many commercial devices.
Figure 9 presents the time evolution of axial-velocity contours
in the fluid. The results show that as the bubbles expand, the
fluid in the chamber achieves higher axial velocity forming
a jet stream. The jet stream subsequently forms a primary
droplet with a long tail. The fluid in the tail is accelerated
back into the primary droplet and a single droplet is formed.
The average velocity of the droplet at a location 0.5 mm from
the nozzle orifice is calculated to be approximately 14 m s1,
which is a typical droplet speed in TIJ devices.

4. Parametric study
This section presents effects of print head geometry,
operational parameters and fluid properties on the device
performance. The geometry of the print head includes nozzle
diameter, heater size and location, chamber size, which can be
1423

A K Sen and J Darabi

900
Heater

t = 2.1 s

Temperature (K)

800

Fluid-film

700
600
500
400

t = 4.5 s

300
0

10

20
30
Time (s)

40

50

Figure 8. Variation of the temperature (K) of the heater and fluid


film with time.
t = 9 s

t = 15 s

t = 22 s

t = 27 s

t = 31 s

Figure 7. Time evolution of the liquid profile along with the


temperature (K) contours inside the liquid.

adjusted during the microfabrication process. The operating


parameters include thermal energy and pulse width, which
can be controlled during operation of the device. The fluid
properties include the density, surface tension and viscosity
that can be tuned by the ink manufacturers. The droplet
volume and speed of the droplets can be precisely modulated
by adjusting the print head geometry, operating conditions
and fluid properties. The droplet volume and speed control
the printing precision and speed. Another design goal is to
minimize the formation of satellite droplets to improve the
print quality. The droplet volume and speed (average) are
calculated by the simulation tool at a distance of 0.5 mm
1424

from the nozzle orifice. The threshold energy is calculated by


determining the minimum thermal energy for which droplet
ejection is observed. The tail length is calculated soon after
the droplet is detached from the chamberfluid interface.
Figure 10 shows effects of the nozzle diameter on droplet
volume and speed. It is observed that there is an increase of the
droplet volume with increase in nozzle diameter. The surface
tension force acting on the fluidair interface is lower for a
larger nozzle, thus more volume of fluid can be ejected for the
same thermal energy. The decrease in the speed of the droplets
may be due to increase in the droplet mass with increase
in the nozzle diameter. Similar trends have been observed
in previous experiments [8]. Variation of droplet volume
and threshold thermal energy with the width of the heater
is depicted in figure 11. The results show that droplet volume
increases with increase in the heater width. Increase in droplet
volume is possibly due to formation of a larger thermal bubble
at a higher heater width, which displaces more fluid from
the chamber. The results also show a decrease in threshold
thermal energy with increase in the heater width which may
be due to a higher rate of heat transfer to the fluid because of
the higher bottom surface area of the heater. The influence of
chamber size and heater location on threshold thermal energy
is presented in figure 12. There is an increase of threshold
thermal energy with increase in the chamber size which may
be due to the increase in static pressure in the chamber for a
larger chamber size. A higher thermal energy is required to
generate a thermal bubble that can generate enough pressure
for the droplet ejection. The results also show increase in the
threshold thermal energy with increase in the distance between
the heater and the nozzle orifice. This is possibly because a
larger amount of heat is transferred (heat loss) to the substrate
which could have been otherwise transferred to the fluid, as
the heater is located far away from the nozzle. The choice of
passivation layer between the heater and the ink chamber is an
important design factor.
Variation of the threshold thermal energy with thermal
conductivity of the passivation layer (for different materials)
is depicted in figure 13. It is observed that there is a significant
decrease of threshold thermal energy with increase in the
thermal conductivity of the passivation layer, as expected.

Droplet ejection performance of a monolithic thermal inkjet print head

t = 4.5 s

15

Droplet speed
Droplet volume

40

14

30

13

20

12

10

11

Droplet speed (m/s)

t = 2.1 s

Droplet volume (pl)a

50

10
5

15

25
35
Nozzle diameter (m)

45

Figure 10. Influence of nozzle diameter on droplet volume and


speed.
10

20

t = 15 s

t = 22 s

Droplet volume (pl)a

T hreshold energy
Droplet diameter

15

6
4

10

2
0

Threshold thermal energy (J)

t = 9 s

5
0

10
15
20
Heater width (m)

25

30

Figure 11. Influence of heater width on droplet volume and


threshold thermal energy.
t = 27 s

Heater location (m)


6
9
12

15

t = 31 s

Figure 9. Axial-velocity contours (m s1) in the fluid.

Four different materials have been evaluatedSiO2, SiN,


SiC, AL and CVD diamond in increasing order of thermal
conductivity.
The influence of the nozzle-guide length on droplet
volume and speed is depicted in figure 14. The results show
that both droplet volume and speed decrease with increase in
the nozzle guide length. This may be due to the fact that
a longer nozzle guide increases the pressure drop or flow
resistance, thus resulting in a smaller and slower droplet.
Decrease in the momentum of the ejected droplet with increase
in the length of the nozzle guide has been reported by [18].
Figure 15 presents influence of thermal energy on the droplet

Threshold energy (J)

30
24
18
12
6

Heater location vs. threshold energy


chamber diameter vs. threshold energy

0
20

40
60
80
Chamber diameter (m)

100

Figure 12. Influence of chamber size and heater location on


threshold thermal energy.

volume and speed. The results show that both droplet volume
and speed increase with increase in the thermal energy. A
higher thermal energy results in faster droplets, thus reducing
the printing time but at the same time the droplet volume
1425

A K Sen and J Darabi

Droplet volume (pl)a

Threshold energy (J)

SiC

12

SiN
AL

CVD diamond

20

Droplet speed

SiO2

Droplet volume

17

14

11

3
0

400

Figure 13. Influence of the thermal conductivity of the passivation


layer on threshold thermal energy.

Droplet volume

12

17

14

11

4
6
Pulse width (s)

10

Density-drop volume
Surface tension- drop volume
Viscosity-drop volume
Density-drop speed
Surface tension-drop speed
Viscosity-droplet speed

8
6

6
9
12
Nozzle guide length (m)

15

Figure 14. Influence of nozzle guide length on droplet volume and


speed.
20

Droplet speed
Droplet volume

12

17

14

11

Droplet speed (m/s)

Droplet volume (pl)a

15

5
4

12
16
Thermal energy(J)

20

Figure 15. Influence of thermal energy on droplet volume and


speed.

increases which tends to reduce the print resolution. The


variation in droplet volume and speed with the change in the
width of the thermal pulse is presented in figure 16. It is
observed that for the same total thermal energy, with increase
in the pulse width there is a slow decrease in the droplet volume

20
18
16
14

12

0.1

5
0

10

Figure 16. Influence of pulse width on droplet volume and speed


for the same total thermal energy.

20

Droplet speed

Droplet speed (m/s)

15

Droplet speed (m/s)

100
200
300
Thermal conductivity (W/mK)

Droplet volume (pl)a

Droplet volume (pl)a

5
0

1426

Droplet speed (m/s)

10

15

1
(mPas)

10
100
(mN/m)

10
1000
10000
(kg/m3)

Figure 17. Influence of density, surface tension and viscosity on


droplet volume and speed.

and increase in the droplet speed. This may be due to decrease


in the heater power (thermal energy per second) for a higher
pulse width (with total thermal energy remaining constant),
which gives rise to a lower rate of heat transfer from the heater
to the fluid. Similar observations are reported by [3]. Effects of
density, surface tension and viscosity of the fluid on the droplet
volume and speed are depicted in figure 17. The results show
that there is a decrease in the droplet volume with increase
in the fluid density. But at the same time higher-density fluid
results in slower droplet, which is not desirable. The results
show that surface tension and the viscosity can significantly
influence the droplet ejection performance. It is observed that
there is a decrease in droplet volume and speed with increase in
surface tension and viscosity. The shape of the ejected droplet
and its tail length is another important performance parameter
of a thermal inkjet device. A higher tail length increases the
chances of formation of potential satellite droplets that can
affect the quality of the inkjet printing process. In the case
of a long tail, the tail breaks into satellite droplets while in
the case of a short tail the tail contracts with the main droplet
eliminating formation of satellite droplet. The effect of ink
properties including density, surface tension and viscosity on
the tail length is illustrated in figure 18. The results indicate
that the tail length decreases with increase in density and

Droplet ejection performance of a monolithic thermal inkjet print head

References

75

Density
Surface tension
Viscosity

Tail length (m) a

60
45
30
15
0
0.1

1
(mPas)

10
100
(mN/m)

1000
10000
(kg/m3)

Figure 18. Influence of ink density, surface tension and viscosity on


tail length.

surface tension and increases slowly with increase in viscosity.


This trend is consistent with the results reported by [19].

5. Conclusion
The droplet ejection performance of a thermally-driven
monolithic inkjet print head based on a dome-shaped ink
chamber, a nozzle guide and a ring-shaped heater was
simulated. The design employs a passivation layer as an
intermediate layer between the ink and the heater, thus
eliminating heater burnout issues and improving reliability.
The design configuration is such that the ink chamber and
the nozzle are inline, thus higher nozzle density can be
achieved. The numerical model is validated by comparing the
model predictions with experimental results for a previously
reported print head design. The agreement between the
model predictions and the experimental results is within
10%, indicating the validity of the model and modeling
methodology. The droplet ejection performance of the
proposed print head design was simulated including the
nucleation and growth of the thermal bubble. A thermal
pulse was used as the driving mechanism for the droplet
ejection. Coupled flow and energy along with bubble
dynamics equations were solved to predict the time evolution
of the fluid interface, droplet volume and speed, heater and
fluid-film temperature and temperature distribution in the ink
present in the chamber. The model was used to investigate
the effects of print head geometry including nozzle diameter,
nozzle-guide length, chamber size, heater location and thermal
conductivity of the passivation layer, operating conditions
including thermal energy and pulse width, and ink properties
on droplet volume and speed. The influence of ink properties
on tail length was also studied. The results of the parametric
study are expected to facilitate the design and operation of the
proposed print head design.

[1] Doane I G 1981 A review of ink-jet printing J. Appl. Photogr.


Eng. 7 1215
[2] Meinhart C D and Zhang H 2000 The flow structure inside a
microfabricated inkjet printhead J. Microelectromech. Syst.
9 6775
[3] Lee J D, Yoon J B, Kim J K, Chung H J, Lee C S, Lee H D,
Lee H J, Kim C K and Han C H 1999 Thermal inkjet
printhead with a monolithically fabricated nozzle plate and
self-aligned ink feed hole J. Microelectromech.
Syst. 8 22936
[4] Beeson R 2000 Desktop inkjet-redefining: the competitive
landscape Proc. 9th Ann. Inkjet Printing Conf. (Scottsdale,
AZ, 412 Oct. 2000) (Kingfield: IMI)
[5] Lee Y S, Shin S J, Kuk K, Kim M S, Shin S and Sohn D K
2004 Lumped modeling of crosstalk behavior of thermal
inkjet print heads Proc. ASME Fluids Engineering Division
2004 vol 260, pp 3918
[6] Tseng F G, Kim C J and Ho C M 2002 A high-resolution
high-frequency monolithic top-shooting microinjector free
of satellite dropspart I: concept, design, and model
J. Microelectromech. Syst. 11 42747
[7] Lee S W, Kim H C, Kuk K and Oh Y S 2002 A monolithic
inkjet print head: dome jet Sensors Actuators A 95 1149
[8] Tseng F G, Kim C J and Ho C M 1998 A novel microinjector
with virtual chamber neck, Micro Electro Mechanical
Systems MEMS 98 Proc.: The 11th Annual International
Workshop (Heidelberg, Germany, Jan 1998) pp 5762
[9] Yan W, Jeffrey B and Lee A 2004 Maskless lithography using
drop-on-demand inkjet printing method Emerging
Lithographic Technologies VIII: Proc. SPIE 5374 62836
[10] Park J H and Oh Y S 2005 Investigation to minimize heater
burnout in thermal thin film print heads Microsystem
Technologies vol 11 (Berlin: Springer) pp 1622
[11] Bae K D, Baek S S, Lim H T, Kuk K and Ro K C 2005
Development of the new thermal inkjet head on SOI wafer
Microelectron. Eng. 7879 15863
[12] Baek S S, Choi B and Oh Y 2004 Design of a high-density
thermal inkjet using heat transfer from CVD diamond
J. Micromech. Microeng. 14 75060
[13] Huang C M, Chen C T, Liou J C, Chiao C C, Wang C Y,
Chuang Y H, Mao C Y and Chen C J 2004 Design and
fabrication of the monolithic inkjet print head 2004 Int.
Conf. on Digital Printing Technologies, (Salt Lake City, UT)
[14] Asai A, Hara T and Endo I 1987 One-dimensional model of
bubble growth and liquid flow in bubble jet printers Japan.
J. Appl. Phys. 26 1794801
[15] Chen P H, Chen W C, Ding P P and Chang S H 1998 Droplet
formation of a thermal side-shooter inkjet printhead Int. J.
Heat Fluid Flow 19 38290
[16] Theofanous T, Biasi L and Isbin H S 1969 A theoretical study
on bubble growth in constant and time-dependent pressure
fields Chem. Eng. Sci. 24 885
[17] Hirt C W and Nichols B D 1981 Volume of fluid (VOF)
method for the dynamics of free boundaries J. Comput.
Phys. 39 20125
[18] Lee Y S, Kim M S, Shin S H and Shin S J 2003 Lumped
modeling of thermal inkjet print head Nanotech
2 48891
[19] Lindemann T, Sassano D, Bellone A, Zengerle R and Coltay P
2004 Three-dimensional CFD simulation of a thermal
bubble jet print head Nanotech 2 22730

1427

You might also like