You are on page 1of 17

Pressure generated at the

rsos.royalsocietypublishing.org
instant of impact between
a liquid droplet and
Research
solid surface
Cite this article: Tatekura Y, Watanabe M, Y. Tatekura1, M. Watanabe1, K. Kobayashi1 and
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

Kobayashi K, Sanada T. 2018 Pressure generated


at the instant of impact between a liquid droplet T. Sanada2
and solid surface. R. Soc. open sci. 5: 181101. 1
Graduate School of Engineering, Hokkaido University, Kita 13, Nishi 8, Kita-ku, Sapporo, Japan
http://dx.doi.org/10.1098/rsos.181101 2
Graduate School of Engineering, Shizuoka University, 3-5-1 Johoku, Naka-ku, Hamamatsu,
Japan
YT, 0000-0003-0417-6233; MW, 0000-0003-0848-3708;
KK, 0000-0001-8842-469X; TS, 0000-0003-4636-8364
Received: 6 July 2018
Accepted: 29 October 2018 The prime objective of this study is to answer the question: How
large is the pressure developed at the instant of a spherical liquid
droplet impact on a solid surface? Engel first proposed that the
maximum pressure rise generated by a spherical liquid
droplet impact on a solid surface is different from the one-
Subject Category:
dimensional water-hammer pressure by a spherical shape
Engineering factor (Engel 1955 J. Res. Natl Bur. Stand. 55(5), 281–298).
Many researchers have since proposed various factors to
Subject Areas: accurately predict the maximum pressure rise. We numerically
mechanical engineering/fluid mechanics/ found that the maximum pressure rise can be predicted by the
computational mechanics combination of water-hammer theory and the shock relation;
then, we analytically extended Engel’s elastic impact model,
by realizing that the progression speed of the contact between
Keywords:
the gas –liquid interface and the solid surface is much
droplet impact, impact pressure, water-hammer
faster than the compression wavefront propagation speed at
pressure, shock wave, shape factor the instant of the impact. We successfully correct Engel’s
theory so that it can accurately provide the maximum pressure
rise at the instant of impact between a spherical liquid droplet
and solid surface, that is, no shape factor appears in the theory.
Author for correspondence:
M. Watanabe
e-mail: masao.watanabe@eng.hokudai.ac.jp
1. Introduction
An understanding of liquid droplet impact onto a rigid solid
surface is needed in a number of technological situations such
as cleaning of surfaces, spray coating, spray cooling and ink-jet
printing. Important practical situations are the erosion of blades
in a steam turbine, in the wet steam erosion of forward-facing
Electronic supplementary material is available components on an aircraft, and erosion of hydraulic machinery
online at https://dx.doi.org/10.6084/m9.figshare. including ship propellers [1–5]. When water drops impact a
c.4302749.

& 2018 The Authors. Published by the Royal Society under the terms of the Creative
Commons Attribution License http://creativecommons.org/licenses/by/4.0/, which permits
unrestricted use, provided the original author and source are credited.
surface above a threshold velocity, they can eventually lead to erosion [6]. Liquid droplet impact is also 2
one of the mechanisms leading to wall thinning, which causes rupture of the piping and unplanned

rsos.royalsocietypublishing.org
outages of pressurized water reactors and boiled water reactors [7]. In order to determine the
threshold velocity for the liquid droplet impact damage, a detailed knowledge of the liquid droplet
impact process is required [4,8]. Obreschkow et al. [9] summarized the most important mechanisms of
erosion that are known to occur during the impact of a liquid drop. A rigorous evaluation of the
impact pressure generated on the solid surface at the instant of impact is essential for studying the
very early stage of liquid droplet impact.
For the accurate prediction of the maximum pressure generated at the instant of impact, a large
number of models have been proposed. Engel [6] estimated the maximum pressure rise DPE from the
atmosphere developed after a spherical liquid drop impact on a solid surface as

R. Soc. open sci. 5: 181101


1
DPE ¼ aE r0 c0 V, (1:1)
2
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

where V is the impact velocity, and r0 and c0 are the density and speed of sound, respectively, in the
undisturbed liquid before the impact. The factor of 12 is due to the ‘so-called’ spherical shape of the
drop. It was found that the coefficient aE is 0.4 using the Schlieren patterns produced by the collision
of a water droplet with a glass plate. Later, Engel [10] proposed a general form of aE. The factor
proposed by Engel became a controversial subject in droplet impact thereafter.
Jenkins & Booker [1] observed the impingement of water droplets, nominally of 2 mm diameter and
roughly of spherical shape on a surface, moving at speeds ranging from 91 to 114 m s21. They found that
their result agrees with the so-called classical water-hammer pressure rise in water DPw, which is based
on the relation

DPw ¼ r0 c0 V: (1:2)

Note that these studies [1,6,10] did not directly measure the pressure; instead, they evaluated the pressure
using the experimentally obtained velocity. Rochester [11] and Rochester & Brunton [12] observed the
impact pressure of a 5.0 mm diameter cylindrical water droplet when it was struck by a bullet at
100 m s21. The pressure rise at instant of the impact is approximately 0.7r0c0V. Further, Rochester
determined that the pressure is reduced at the centre of impact because the surface or the bullet
deforms slightly during impact. The implication of this is that the pressure rises in the very early
stage of impact against a solid surface are approximately 1.4 times those measured in the experiments.
Therefore, he concluded that in an impact against a solid surface, these pressures become r0c0V.
Bourne [13] evaluated the impact response of polymethylmethacrylate (PMMA) by measuring
pressure in the region beneath the impacting liquid droplet with a radius of 2 mm by using
polyvinylidene fluoride gauges at an impact velocity of 600 m s21. He found that the initial impact
pressure was approximately 0.8 GPa, which was just below the Hugoniot elastic limit for PMMA.
Heymann [14] introduced a two-dimensional model that represents a thin parallel slice taken
vertically through the impacting drop. He inferred that the pressure rises are virtually uniform and
equal to

DPs ¼ r0 sV, (1:3)

at the first instant of contact for the impact of a liquid droplet, where s in equation (1.3) is the compression
wavefront speed [15] with a pressure that increases by DPs with respect to the one in the undisturbed
liquid pressure. His inference that the pressure rise at the instant of the impact for a liquid drop can
be predicted by equation (1.3) is far from obvious. It is this inference that we primarily discuss in
detail throughout this paper. Lesser [16] also claimed that the impact pressure at the instant of the
impact on the centre can be estimated from the shock relation. Later, Field [17] explained that for the
impact of a liquid droplet on a solid surface, the pressure rise on the central axis at the instant of
impact is estimated by equation (1.3). Korobkin [18,19] generalized Lesser’s result to the whole region
behind the shock front in his analysis on the impact by a flat plate on a circular liquid drop.
The numerical analysis conducted by Hwang & Hammitt [20] explored a new aspect of the study of
liquid droplet impact. They investigated the impact process of cylindrical, spherical and conical water
droplets on a solid plane surface using numerical methods. They found that the impact pressure rise
reaches a peak of 0.7r0c0V. Haller [21] and Haller et al. [22] also numerically investigated the fluid
dynamics of high speed (500 m s21) small size (200 mm in diameter) droplet impact on a solid surface
using a high-resolution axisymmetric solver for the Euler equations. Their numerical results support
Lesser’s assumption.
These studies show that extensive efforts have been made toward the understanding of the pressure 3
generated at droplet impact; however, the maximum pressure rise at the impact on the central axis has yet

rsos.royalsocietypublishing.org
to be indisputably determined. Recent various numerical analyses indeed provide a variety of results.
Sanada et al. [23] carried out a numerical simulation of an axisymmetric spherical liquid droplet
impact on a solid surface, using the ghost fluid method [24]. They found that the maximum pressure
rise of the point closest to the solid surface can be well predicted by equation (1.2). Xiong et al. [7]
found that the maximum average impact pressure given by their simulation agrees well with the
Heymann correlation pressures [14] using the moving particle semi-implicit method. Sanada et al. [25]
and Kondo & Ando [26] numerically studied axisymmetric and cylindrical droplet impact on a solid
surface, respectively, using the shock-interface capturing scheme [27]. Their results support Engel [6],
i.e. equation (1.1). Hsu et al. [28] also supported equation (1.1) as an empirical formulation that can
predict the maximum pressure rise for the impingement of a spherical liquid drop.

R. Soc. open sci. 5: 181101


Note that pressure impulse theory, which Cooker & Peregrine [29] and Philippi et al. [30] used, and
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

most of classical work on the fluid–solid impact, including the seminal paper by Wagner [31], Howison
et al. [32] and Scolan & Korobkin [33], were carried out under the assumption of an ideal and
incompressible liquid. However, the compressibility of the liquid plays the most significant role in the
prediction of the maximum pressure rise. The most general methods to investigate the pressure
generated at the instant of impact between a liquid droplet and solid surface are those that
appropriately take the liquid compressibility into consideration.
The prime objective of this study is to answer the question: How large is the pressure developed at the instant
of a liquid droplet impact on a solid surface? by shedding light on the shape factor that is still controversial. To
answer this question, we consider the impact of a spherical liquid droplet on a solid surface under the
assumption that liquid droplet keeps its spherical shape at the instant of the impact, although it is well
known that a liquid droplet impacting on a solid surface deforms before the liquid droplet makes contact
with the surface [34 –36] or entraps a small air bubble under its centre under atmospheric conditions
[37,38]. The reason is that we primarily discuss the role of the spherical shape factor introduced by Engel
[6], who also used the assumption that the liquid droplet keeps its spherical shape at the instant of the
impact, in the central impact pressure rise. We numerically estimate the central impact pressure (defined
here as the pressure developed at the instant of the impact on the contact point on the axis of symmetry)
by solving two-dimensional axisymmetric compressible Euler equations. We then extend Engel’s elastic
impact model [6] taking the limit of the time from impact for the pressure to reach its maximum to 0, by
realizing that the progression speed of the contact between the gas –liquid interface and the solid surface
is much faster than the compression wavefront propagation speed at the instant of the impact. We
correct Engel’s elastic impact model to find that the factor aE/2 appearing in equation (1.1) converges to
1 at the instant of the impact.

2. Previous models for one-dimensional analyses of impact pressure rise


As stated in the previous section, the water-hammer theory has been often used in the analysis of liquid
droplet impact pressure rise. We also contrast the liquid droplet impact pressure with the water-hammer
pressure rise in later sections; hence, in this section, we briefly review a couple of previous
one-dimensional models used in the estimation of impact pressure.
A sudden halt of flow in a pipe is considered to be equivalent to a liquid droplet impact on a solid
surface. Cook [39] was one of the first to realize that the pressure generated on an element of a surface
at its first encounter with water moving at a finite velocity would be given by the water-hammer
pressure rise. Gardner [40] and de Haller [41] extended this analysis by accounting for the
compressibility of the solid.
Suppose a uniform flow of a liquid moving with a speed of V in a pipe to the right in figure 1a. When
the pipe is suddenly closed at the valve, a high pressure builds and propagates upstream (to the left in
figure 1a) with a speed of (s 2 V ), where s is the compression wavefront speed in a quiescent liquid.
Figure 1a can be redrawn as figure 1b using the coordinate system moving with a speed of (s 2 V ).
Heymann [14], on the other hand, used the so-called shock relation to estimate the pressure rise
developed at a liquid droplet on a solid surface. He considered that a moving solid plate with a
velocity V collides with a stationary liquid droplet. Then, a shock front is generated with a speed of s
moving into a quiescent medium, as shown in figure 2a. The liquid velocity induced behind the shock
front is the same as the plate velocity V. Figure 2a can be redrawn as figure 2b using the coordinate
system moving with a velocity of s.
(a) r1, p1 (b) r1, p1 4

rsos.royalsocietypublishing.org
V0 = V V1 = 0
s s–V
s–V
r0, p0 r0, p0

Figure 1. Water-hammer pressure rise in (a) a stationary coordinate system, and (b) a moving coordinate system with a speed of
(s 2 V ), where s is the compression wavefront speed in a quiescent liquid and r0, p0, and V are density, pressure and speed of the
uniform flow of the undisturbed liquid, respectively. A high pressure builds and propagates upstream (to the left in (a)) with a
speed of (s 2 V ). r1 and p1 are the density and pressure of the liquid that the pressure wave has reached, respectively.

R. Soc. open sci. 5: 181101


r1, p1
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

(a) r1, p1 (b)

V0 = 0 V1 = –V
s s–V
–s
r0, p0 r0, p0

Figure 2. Shock relation in (a) a stationary coordinate system, and (b) a moving coordinate system with the compression wavefront
speed s. The shock front is generated with a speed of s moving into a quiescent medium with density r0 and pressure p0 as shown
in (a), where r1 and p1 are the fluid density and pressure of the liquid behind the compression wavefront, respectively.

Note that both figures 1b and 2b are identical; hence, the conservation equations of mass, momentum
and energy for both cases should be identical and are, respectively, written as
r0 s ¼ r1 (s  V), (2:1)
2 2
p0 þ r0 s ¼ p1 þ r1 (s  V) ¼ p1 þ r0 s(s  V) (2:2)

and
2  " #
s (s  V)2
p0 s þ r0 s þ e0 ¼ p1 (s  V) þ r1 (s  V) þ e1 : (2:3)
2 2

Equations (2.1) and (2.2) lead to the following expression of pressure rise DPs:
DPs ¼ p1  p0 ¼ r0 sV: (1:3)

The classical water-hammer pressure rise (1.2) can be obtained by simplifying equation (1.3) using the
commonly used assumption that the compression wavefront speed can be replaced with c0, the speed
of sound in an undisturbed liquid:
s ¼ c0 : (2:4)

Further review on this topic can be found in Brunton & Rochester [42] and Lesser & Field [43].
We now consider s as a variable instead of a constant to more accurately predict the maximum
pressure rise. We need an additional equation to close the equation set. We examine two different
equations for properly evaluating s. The first equation was empirically developed by Heymann [15]
and is the approximate shock velocity –particle velocity relationship:
  
V
s ¼ c0 1 þ k ¼ c0 (1 þ kM0 ), (2:5)
c0

where
V
M0 ¼ (2:6)
c0
is the ‘impact Mach number’ and k is a dimensionless constant. Heymann [14] obtained the impact
pressure rise using the approximate shock velocity –particle velocity relationship DPH:
DPH ¼ r0 c0 V(1 þ kM0 ): (2:7)
Heymann [15] found that the value k ¼ 2 fits the data for water up to about M0 ¼ 1.2. Substituting the 5
values c0 ¼ 1497 m s21 and V ¼ 304.8 m s21 used by Engel [44] into equation (2.7), an impact pressure

rsos.royalsocietypublishing.org
of 1.41r0c0V can be obtained with the assumption of a rigid solid impact plate.
The second equation is the stiffened-gas equation of state proposed by Harlow & Amsden [45]:
p ¼ (g  1)re  gP, (2:8)

where e is the internal energy per unit mass, g and P are constants ( g ¼ 1.4 and P ¼ 0 for air; g ¼ 4.4 and
P ¼ 6.0  108 for water). Equation (2.8) is widely used for computational analyses to investigate
compressible gas –liquid two-phase flow [46–48]. Saurel & Abgrall [46] reported that it is possible to
describe compressible liquid under high pressure with reasonable accuracy using the stiffened-gas
equation of state, and that each phase of the flows involving various materials may be described by
an equation of state of this type. We also implemented equation (2.8) in our numerical analysis, as

R. Soc. open sci. 5: 181101


explained in the next section.
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

Equations (2.1), (2.2) and (2.3) with the stiffened-gas equation of state (2.8) give the speed of sound for
the stiffened-gas [49]:
2 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi3
(g þ 1)M 0 (g þ 1)2 M20 5
s ¼ co 4 þ 1þ : (2:9)
4 16

Using equation (2.9), we obtain the impact pressure rise with the use of stiffened-gas equation of
state DPsg:
2 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi3
(g þ 1)M 0 (g þ 1)2 M20 5
DPsg ¼ ro co V 4 þ 1þ : (2:10)
4 16

We found in this section that the one-dimensional liquid impact pressure rise DPs on a solid surface can
be formally written
DPs ¼ r0 sV, (2:11)

regardless of the choice of the model. All these models were derived from the same conservation
equations; however, different equations for the compression wavefront speeds were used. We refer to
these pressure rises, DPw: equation (1.2), DPH: equation (2.7) and DPsg: equation (2.10), as the classical
water-hammer impact pressure rise, Heymann impact pressure rise and stiffened-gas impact pressure
rise, respectively, in the following discussion.

3. Numerical evaluation of the pressure developed immediately after the


liquid droplet impact
3.1. Numerical method
We numerically investigated the liquid droplet impact pressure rise on a solid surface. We analysed the
pressure generation at droplet impact by solving the two-dimensional axisymmetric compressible Euler
equations [22,50]. To close the Euler equations, we implemented the stiffened-gas equation of state (2.8)
in our numerical analysis; hence, we compare the pressure rise due to the liquid droplet impact obtained
by numerical analysis with the stiffened-gas pressure rise DPsg (2.10). We solved the flow field using the
level-set method [51,52] combined with the ghost fluid method [24,53,54]. We used a finite difference
method for discretization of the system. The third-order TVD Runge –Kutta scheme was employed to
march the equations forward in time, and the third-order ENO-LLF scheme was employed to
calculate the convection term [55,56]. The third-order TVD Runge –Kutta scheme and fifth order
WENO scheme [57] was employed to advance the level-set function c, where sets of c ¼ 0, c , 0 and
c . 0 represent the gas –liquid interface, one fluid and the other fluid, respectively. We calculated the
physical properties only in the liquid phase with constant gas pressure. Haller et al. [22] carried out a
numerical simulation of droplet impact on a solid surface by solving flow fields of both liquid and
gas phases. Their results succeeded to reproduce the theoretical results by Lesser [16] who did not
consider the dynamics of the gas phase. Hence, we can neglect the contribution of gas that was also
neglected by the theoretical models for the purpose of this study.
We introduced the assumption that the portion of the droplet that has not collided remains spherical.
This should hold in the computation partly because we mainly investigate the pressure development
6
(a) calculation domain (b)

rsos.royalsocietypublishing.org
droplet
droplet

z z

R. Soc. open sci. 5: 181101


(c)
solid surface
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

Cell-1 Cell-2 Cell-3 Cell-4 Cell-5


+ + + + + r
r
O

Figure 3. Calculation domain. The square shown in (a) is the computational domain used in this study. This computational domain
is a portion of the configuration of the liquid droplet impact on the solid surface shown in (b). The length of the side of the square
is R/4, where R is the droplet radius at the initial state and was set to 1 mm. The size of the square calculation cell Dr is non-
dimensionalized: Dr ¼ Dr=R with the range of 215  Dr  24 . The initial pressure was set to 0.1013 MPa. The initial
density and speed of sound were set to 1000 kg m23 and 1625 m s21 for the liquid.

immediately after the droplet impact, where the deformation may be insignificant, and partly because we
discuss the legitimacy of the theoretical models by comparing them with the numerical results. Therefore,
the contour line of c ¼ 0 is enforced to maintain a circle whose centre moves with impact velocity V
in z . 0.
Impact velocity V was set to 100 m s21 because this order of magnitude of impact velocity can be
typically observed in cleaning technologies for semiconductor device processes [23,50,58,59], in single
water droplet impact experiments [11,60,61] and in numerical simulations of liquid droplet impact
[9,26]. Note that the difference between c0 and s is hardly noticeable if the value of V is much smaller,
and the stiffened-gas equation of state (2.8) is not applicable if the value of V is much larger.
We calculated the pressure of the fluid in the cell contacting the solid surface. We then identified this
pressure using the pressure on the solid surface. We refer to these cells adjacent to the solid surface as the
surface cells. We define Cell-1 to be the left-most surface cell contacting with the z-axis, as shown in
figure 3c. We consider the gas –liquid interface to impact on the solid surface when the fluid in the
surface cell that is adjacent to the point turns from gas to liquid, i.e. when the interface crosses
the centre of the surface cell. We also consider the droplet to impact on the solid wall when the
gas –liquid interface impacts on the solid surface in Cell-1. The central pressure is defined as the
pressure in Cell-1. We discuss the temporal change of the central pressure after the droplet impacts on
the solid surface. We also investigated the pressures in other surface cells, particularly Cell-2, Cell-3,
Cell-4 and Cell-5, as shown in figure 3c.

3.2. Estimation of the central pressure


We now examine the maximum central pressure rise DPmax at the instant of impact. Figure 4 shows
the temporal changes of the pressure in Cell-1 calculated with various cell sizes, where the non-
dimensionalized time t was defined by t ¼ Vt=R. We use the following expression for the maximum
central pressure rise:
DPmax ¼ aDPsg ¼ ar0 sV, (3:1)

where a is defined as the ratio of the calculated maximum central pressure rise to the stiffened-gas impact
pressure rise, and s is defined by equation (2.9). Throughout this paper, we use the constant speed of
compression wavefront s ¼ 1760 m s21, which is obtained by substituting V ¼ 100 m s21 and c0 ¼
1625 m s21 in equation (2.9). It is a remarkable finding that a depends on Dr ; a reaches 0.994 in
the case of Dr ¼ 215 , while it is 0.455 in the case of Dr ¼ 24 , as shown in figure 4. Surprisingly, the
variation in the value of a is more than double.
(a) 1.2 (b) 1.2 7
: Dr* = 2–4
: Dr* = 2–5

rsos.royalsocietypublishing.org
1.0 1.0
: Dr* = 2–6
: Dr* = 2–7
0.8 : Dr* = 2–8 0.8
: Dr* = 2–9

p/r0sV
p/r0sV

0.6 0.6
: Dr* = 2–10
: Dr* = 2–11
0.4 0.4 : Dr* = 2–12
: Dr* = 2–13
0.2 0.2 : Dr* = 2–14
: Dr* = 2–15

R. Soc. open sci. 5: 181101


0 1 2 3 4 5 0 0.4 0.8 1.2 1.6 2.0
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

t* × 103 t* × 104
Figure 4. Convergence of the maximum central pressure obtained by numerical analysis as the non-dimensional calculation cell size
Dr  decreases: (a) 29  Dr   24 and (b) 215  Dr   210 . The calculated pressure rise was normalized by the
stiffened-gas impact pressure rise DPsg defined by equation (2.10). Note that the scale of the abscissa of (b) is 25 times
greater than that of (a).

(a) 1 (b) 10–2

10–3
10–1
te* 10–4
1–a

10–2 te* = 0.0457 Dr*0.833


10–5

10–3 10–6
10 102 103 104 105 10 102 103 104 105
1/Dr* 1/Dr*
Figure 5. (a) The dependency of a (the ratio of the calculated maximum central pressure rise to the the stiffened-gas impact
pressure rise as in equation (2.10)) on Dr . (1 2 a) (in log-scale) is plotted on the ordinate and Dr (in log-scale) is
plotted on the abscissa. (b) The dependency of t (non-dimensional elapsed time from the instant of impact until the
pressure reaches the maximum) on Dr  . t (in log-scale) is plotted on the ordinate and Dr  (in log-scale) is plotted on
the abscissa The regression line shows that te is approximately proportional to Dr .

We evaluate a for each Dr . Figure 5a shows that a monotonically converges to 1 when Dr decreases.
Therefore, we conclude that the maximum central pressure rise at the instant of impact DPmax is given by

DPmax ¼ DPsg ¼ r0 sV: (3:2)

Note that the results of the numerical simulation provide the maximum central pressure rise that can
be expressed by equation (3.2) under very restricted conditions, where Dr is sufficiently small. For
instance, we should choose a Dr that is less than 2214 to obtain an a greater than 0.99. We found that
significantly small calculation cells are required to resolve the magnitude of the pressure at the instant
of impact.
We also found in figure 4 that te, the elapsed time for the pressure to reach its maximum from the
instant of impact, decreases as Dr decreases. Non-dimensionalized elapsed time te ( ¼ te V=R) is
plotted in figure 5b. Note that the impact pressure attains the maximum pressure because of
the elastic compression of the fluid in the cell; hence, te decreases as Dr decreases. Therefore, in
the limit taken as Dr approaches 0, the elapsed time to reach the maximum central pressure
converges to 0.
(a) 1.2 (b) 1.2 8
: Cell-1

rsos.royalsocietypublishing.org
1.0 : Cell-2 1.0
: Cell-3
0.8 : Cell-4 0.8
: Cell-5
p/r0sV

p/r0sV
0.6 0.6
: Cell-1
0.4 0.4 : Cell-2
: Cell-3
0.2 0.2 : Cell-4
: Cell-5

R. Soc. open sci. 5: 181101


0 0.5 1.0 1.5 2.0 2.5 0 0.4 0.8 1.2 1.6
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

t* × 102 t* × 103

(c) 1.2

1.0

0.8
p/r0sV

0.6
: Cell-1
: Cell-2
0.4
: Cell-3
: Cell-4
0.2 : Cell-5

0 0.5 1.0 1.5 2.0


t* × 104
Figure 6. Pressure developments in the first five surface cells for the cases of (a) Dr ¼ 24 , (b) Dr ¼ 27 and (c)
Dr  ¼ 210 . (a) The pressure in Cell-2 begins to rise only after the pressure in Cell-1 has reached the maximum; then, a is
0.455. (b) The pressure in Cell-2 begins to rise when the pressure in Cell-1 is still increasing; then, a is much greater than
0.455 but much less than 1.0. (c) The pressure in Cell-2 begins to rise immediately after the pressure in Cell-1 has begun to
rise; then, a is very close to 1.

3.3. Maximum central pressure dependency on calculation cell size


We found that a can converge to 1 as long as Dr is sufficiently small in §3.2; however, a can reach at most
0.455 when Dr ¼ 24 : We now discuss the dependency of a on Dr . For a large cell size (Dr ¼ 24 in
figure 6a), the pressure in Cell-2 begins to rise only after the pressure in Cell-1 reaches the maximum;
that is, Cell-2, Cell-3, Cell-4 and Cell-5 are filled with gas while the pressure in Cell-1 is increasing;
then, a is 0.455. For a medium cell size (Dr ¼ 27 in figure 6b), the pressure in Cell-2 begins to rise
when the pressure in Cell-1 is still increasing; that is, the fluid in Cell-2 turns from gas to liquid while
the pressure in Cell-1 increases; then, a is much greater than 0.455 but much less than 1.0. For a small
cell size (Dr ¼ 210 in figure 6c), the pressure in Cell-2 begins to rise immediately after the pressure in
Cell-1 has begun to rise; that is, Cell-2 is almost filled with liquid while the pressure in Cell-1 is
increasing; then, a is very close to 1. These results suggest that the status of Cell-2, i.e. either liquid or
gas, should greatly affect the central maximum pressure.
We now introduce the ratio of time f:

t2
f¼ , (3:3)
t1

where t1 is the non-dimensional time which it takes for the pressure in Cell-1 to rise from p0 to Pmax, and
t2 is the non-dimensional duration over which the fluid in Cell-2 is gas for t1 . Figure 7 shows that
a monotonically increases as f decreases. We can see that the time duration for which the fluid in
Cell-2 is liquid while the pressure in Cell-1 is increasing totally dominates the value of a.
1 9

rsos.royalsocietypublishing.org
10–1

1–a
10–2

R. Soc. open sci. 5: 181101


10–3
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

10–4 10–3 10–2 10–1 1


f
Figure 7. The ratio of time f, f ¼ t2 =t1 , is plotted on the abscissa (in log-scale) with (1 2 a) on the ordinate (in log-scale).
The values of f for Dr of 224, 227 and 2210 are 1.00, 7.86  1022 and 6.46  1023, respectively. Here, a monotonically
increases as f decreases. This indicates that the time duration for which the fluid in Cell-2 is liquid while the pressure in Cell-1 is
increasing completely dominates the value of a.

When the fluid in Cell-2 turns to liquid immediately after the pressure rises in Cell-1, the liquid in
Cell-1 is confined to the space bounded by the liquid in Cell-2, for which the pressure is about the
same as that of Cell-1; hence, the liquid in Cell-1 can be compressed primarily in the direction normal
to the solid surface. This compression can be thought of as the one-dimensional liquid compression
by which the pressure generated can be predicted well by equation (2.10). Therefore, a is essentially
equal to 1 when f is sufficiently close to 0.
When the fluid in Cell-2 remains gas after the pressure rises in Cell-1, the liquid in Cell-1 is confined
to the space bounded by the gas in Cell-2; hence, the liquid in Cell-1 is able not only to be compressed in
the normal direction but also to be expanded in the direction parallel to the solid surface. In other words,
the liquid in Cell-1 can flow in the direction parallel to the solid surface. As Bowden & Field [2]
explained, a cannot reach 1 when the liquid in Cell-1 can flow. According to them, a can reach 1
when the liquid in Cell-1 cannot flow, that is, the flow can be assumed to be one-dimensional.
Therefore, a is significantly smaller than 1 when f is close to 1.

3.4. Impact of an interface with a finite curvature


According to the numerical analyses conducted by Sanada et al. [25] and Kondo & Ando [26], the
pressure rise in liquid impact on a solid wall can be well predicted by using the Heymann impact
pressure rise (equation (2.7)) when the impact can be considered as a one-dimensional impact, where
the radius of curvature of the impacting interface is infinite, i.e. a plane interface. However, they also
found that liquid droplet impact pressure rise on the same solid wall can be significantly reduced
when the radius of curvature of the impacting interface is finite: they numerically reproduced a
cylindrical droplet impact on a solid wall. The latter pressure was approximately 0.2 of the former
pressure. Their results apparently supported Engel [6], i.e. equation (1.1). We now examine whether
the liquid impact pressure rise can be reduced by the stiffened-gas pressure rise (equation (2.10))
when the interface that impacts on a solid surface has a finite radius of curvature using the numerical
analysis.
We found that the impact pressure rise can reach the one-dimensional liquid impact pressure rise
when Dr is sufficiently small in our numerical simulation. We must carefully examine the possibility
that the interface of the liquid droplet might be plane with an infinite radius of curvature when the
liquid droplet impacts on a solid surface for sufficiently small Dr because the plane interface impact
might be equivalent to a one-dimensional impact, which leads to the stiffened-gas pressure rise
(equation (2.10)). Therefore, we need to confirm whether the liquid droplet impacts a solid surface
with a finite radius of curvature.
When a gas –liquid interface impacting on a solid surface is a plane, the impacts of the interface
simultaneously occur in adjacent multiple cells; hence, the impacts of the interface should not
(b) Dr* 10
(a)
z

rsos.royalsocietypublishing.org
4 × 10–4Dr*
droplet surface
(c)

1 2 3 4 5
r (d)
O

Figure 8. Impact of an interface with a finite curvature with a solid surface. No simultaneous interface impact occurs in any adjacent

R. Soc. open sci. 5: 181101


cells. (a) Location of the interface in Cell-1, Cell-2, Cell-3, Cell-4, and Cell-5, (b) t  =Dt  ¼ 0, (c) t  =Dt  ¼ 1, (d) t  =Dt  ¼ 3.
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

Note that the ordinates in (b – d) are greatly magnified with respect to (a). The numerical calculation was carried out under the
conditions that Dr  ¼ 215 and Dt  ¼ 3:73  1010 .

simultaneously occur in adjacent cells in figure 8 when the interface has a finite radius of curvature.
Figure 9a shows the pressure development at the five surface cells in figure 8, although the pressures
observed at different surface cells are identical. Figure 9b shows an enlarged view of the very early
stage of droplet impact. Therefore, we conclude that we successfully reproduced the impact of a
droplet which has a finite radius curvature with a solid surface.
We now consider the maximum central pressure rise. Figure 9a shows that the maximum pressure
rises observed at the surface cells Cell-1 to Cell-5 shown in figure 8 are identical. We find that the
factor a defined by equation (3.1) is 0.994 in figure 9. Therefore, we conclude that the maximum
central pressure rise that develops when a droplet impacts a solid surface can be accurately predicted
by equation (2.10), even with a finite radius of curvature interface.
To close this section, we add the following as the final remarks. It requires extremely high resolutions
in both the spatial and temporal domains to obtain the numerical results that confirm that the maximum
central pressure rise can be predicted by equation (2.10). We are convinced that insufficient resolutions in
either the numerical or experimental studies have caused the confusion in determining the factor.

4. Equation for the central pressure resulting from a liquid droplet impact
We now return to the Engel pressure rise DPE:
1
DPE ¼ aE r0 c0 V: (1:1)
2
Engel stated with regard to equation (1.1) that this is the well-known water-hammer equation multiplied by
factor aE/2, which comes from the spherical shape of the liquid droplet; indeed, it has long been understood
that the factor of 12 is considered to be a consequence of the spherical shape of the drop [28]. However,
our numerical analysis showed that the impact pressure rise due to the spherical liquid droplet can be
predicted by equation (2.10), where no shape factor is required. In this section, we reconsider the
origin of the factor of 12 and then discuss the legitimacy of using equation (1.1) for predicting
the maximum central pressure rise.

4.1. Maximum pressure rise proposed by Engel


Engel [6] derived equation (1.1) for the maximum pressure rise DPE that develops at the time Dt after a
liquid droplet impacts on a solid surface. We now briefly introduce the analysis by Engel. She considered
that the liquid droplet impact consists of two sequential (not simultaneous) processes: a compression
which is an instantaneous process and a subsequent displacement which takes a finite time Dt, as
shown in figure 10 drawn based on fig. 14 in [6]. Engel introduced a coefficient aE, which is an index
of the fraction of velocity V that is imparted to the liquid molecules on average. She assumed that the
coefficient aE must be governed mainly by the extent of divergence of the compression wave as it
spreads through the spherical droplet. She also assumed that the amount of divergence of the wave
decreases and the value of the coefficient aE approaches unity as the impact velocity increases.
(a) 1.2 (b) 0.020 11

rsos.royalsocietypublishing.org
1.0
0.016 : Cell-1
: Cell-2
0.8 : Cell-3
0.012 : Cell-4
p/r0sV

p/r0sV
0.6 : Cell-5
: Cell-1
: Cell-2 0.008
0.4
: Cell-3
: Cell-4 0.004
0.2 : Cell-5

R. Soc. open sci. 5: 181101


0 100 200 300 400 500 0 2 4 6 8 10
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

t*/Dt* t*/Dt*
Figure 9. Pressure development at the first five surface cells at droplet impact: (a) pressure development; (b) enlarged view of (a) at
a very early stage of the impact. The interface impacts on the solid surface in Cell-1 at t  =Dt  ¼ 1. The pressure of the liquid in
Cell-1 then rises while the pressure of the liquid in the other cells remains unchanged. The shape of the interface at t  =Dt  ¼ 1 is
drawn in figure 8c. The interface impacts on the solid surface in Cell-2 at t  =Dt  ¼ 3. The shape of the interface at t  =Dt  ¼ 3
is drawn in figure 8d. The interface impacts on the solid surface in Cell-3, Cell-4 and Cell-5 at t  =Dt  ¼ 4, 6 and 7, respectively.

rwd

l-plane

d-plane
wd
d
O r

Figure 10. Cross section of the liquid droplet reproduced from fig. 14 in [6]. First, the impact of a droplet of radius R (dashed curve)
with a solid plane surface forces the radius of curvature of the droplet in the vicinity of the impact to change instantaneously to R0
(solid curve), and a compression wavefront to propagate through the liquid droplet, which induces a velocity in the z-direction in the
liquid (the compression process). Second, the moving solid plane surface of velocity V travels through the liquid whose average
velocity may be written as aEV in the time interval Dt to reach the d-plane. The relative velocity of the moving solid plane
surface to the liquid droplet therefore becomes (1 2 aE)V; hence, the compressed cap, which is shown as the light grey
region, with height d, d ¼ (1 2 aE)VDt, is displaced into a very thin cylinder of liquid that is in radial flow (the
displacement process). The liquid cylinder with a base radius of rvd formed between the d-plane and the l-plane has the
same volume as that of the spherical cap whose base is on the l-plane with a height of vd.

To obtain the maximum pressure rise DP at time t ¼ Dt, Engel considered the total mass of liquid m
that gains momentum during a finite time interval Dt, and consequently all the force that acts to produce
the momentum over time interval Dt. Engel considered a separated pair of regions. First, she considered
the mass m0 of the spherical cap, which gains a momentum that causes a radial flow. She defined the
l-plane to be the upper boundary of the radial flow and considered it to be an effective striking
surface with which compression wavelets are initiated at the points of contact of the remaining liquid
droplet. The distance between the d- and l-planes is (v 2 1)d where v  1. She noted that in the limit,
the volume of a spherical cap is one-half the volume of a cylinder that has the same height and base area.
Engel then assumed that the first liquid of the droplet that encounters the moving solid plane surface 12
during the time interval Dt is displaced to form a thin cylinder the height of which is determined so that

rsos.royalsocietypublishing.org
the volume of the spherical cap below the l-plane before the displacement (the union of medium grey
and light grey regions in figure 10) is equal to that of the cylinder (the union of medium grey and
dark grey regions in figure 10); hence, v is equal to 2. During time Dt, which is the time needed for
the moving solid plane surface with the relative velocity (1 2 aE)V to move a distance d through the
droplet, the l-plane moves a distance vd. Velocity Vl of the l-plane is then given by

vd
Vl ¼ ¼ v(1  aE )V: (4:1)
Dt
The mass of the spherical cap (the union of medium grey and light grey regions in figure 10) whose base
radius is rvd, is given by

R. Soc. open sci. 5: 181101


vp 2
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

m0 ¼ r0 vm0 ¼ r d, (4:2)
2 vd
where vm0 is the volume of the spherical cap. Engel mentioned that although the mass of liquid m0 moves
horizontally after time Dt, it was given as a velocity in the collision direction z during this time interval
because it was traversed by the compression wave.
Engel next considered the region (with mass m00 ) of the liquid droplet that has been traversed by the
compression wave during the time interval Dt, which is shown as a hatched region in figure 10; then,
she assumed that the average velocity of the liquid in this region may be written as aEV. Suppose that
the radius of the circular intersection between the l-plane and the surface of the spherical cap is r.
The compression wavelet can be generated on this circular intersection with 0  z  vd when the
l-plane hits the surface of the spherical cap as shown in figure 10. The l-plane is relatively stationary
with respect to the droplet in the compression process because it keeps contact with a point at the
first instant of impact, while it moves a distance vd in the displacement process over time Dt. The
compression wave consists of each compression wavelet, which travels with the velocity s. Engel
assumed that the compression wavelet travels only in the z-direction. Recall that z ¼ vd at r ¼ rvd; hence,

r2vd ¼ 2R0 vd  (vd)2 ≃ 2R0 vd: (4:3)

Let z0 be the distance that the compression wavelet that originated from the circular intersection with
radius r travels above the l-plane at z ¼ vd. The distance Dz between the circular intersection and the
l-plane can be written as
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Dz ¼ vd  z ¼ vd  R þ R2  r2 : (4:4)

Suppose that the l-plane moves by Dz to reach z ¼ vd within time Dt after the compression wavelet was
generated:

Dz
Dt ¼ , (4:5)
v(1  aE )V

then,
   
s s
z0 ¼ Dts  Dz ¼  1 Dz ¼  1 Dz ¼ bDz, (4:6)
v(1  aE )V Vl

for Dz, 0  Dz  2d, where b is defined as b ¼ s=Vl  1. Note that the locus of points z0 gives the
envelope of the compression wavelets.
Next, Engel evaluated the volume of the region vm00 that has been traversed by the compression wave
during the time interval Dt, which is indicated by the hatched region in figure 10. Volume vm00 is that of
the body of revolution bounded by the envelope and l-plane at z ¼ vd. Then,
ð vbd  vbd
z03 (vd  R) 02
vm00 ¼ p r2 dz0 ¼  2 þ z  (v2 d2  2vdR)z0
0 3b b z0 ¼0
 v  v
¼ v bpd R  d ≃ bprvd d:
2 2 2
(4:7)
3 2

Engel estimated that the average rate of change of momentum of liquid in volume (vm0 þ vm00 ) is (m0 þ
m00 )aEV/Dt. She also assumed that all of the force that has acted to produce the liquid momentum is
written as pr2vd DPE . She finally obtained the following: 13
0 00
(m þ m )aE V r0 (vm0 þ vm00 )aE V aE

rsos.royalsocietypublishing.org
DPE ¼ ¼ ¼ r sV: (4:8)
pr2vd Dt pr2vd Dt 2 0

4.2. Instantaneous maximum pressure at the instant of liquid droplet impact


The most significant aspect of the analysis by Engel is that she assumed that Dt is finite; however, the results
obtained by numerical analysis in §3 show that the time elapsed for the pressure on the central axis to reach
the maximum pressure converges to 0. Therefore, we should examine whether equation (4.8) is valid even
for the instantaneous maximum pressure rise. In other words, we should answer the question: Is equation
(4.8) valid in the limit of Dt to 0, although Dt does not explicitly appear in equation (4.8)?

R. Soc. open sci. 5: 181101


First, we discuss the compression process. Engel assumed that the radius of curvature of the droplet
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

in the vicinity of the impact changes instantaneously with the point impact remaining in the compression
process, as shown in figure 10. We should understand that it takes a finite compression time Dt0 for
the liquid in the vicinity of the impact to be compressed, where Dt0 is the time that it takes for the
compression wavefront to travel through the compressed region. We now reexamine how the droplet
can be compressed in Dt0 .
Our numerical results in §3.4 showed that while the intersection advances a distance of 5Dr ( ¼ 5  215 )
in the r-direction during this compression time of 6Dt ( ¼ 2:28  109 ), the solid surface moves by a
distance 7:32  105 Dr and the compression wavefront travels in the liquid droplet by a distance
1:30  103 Dr in the z-direction. We are hence considering an extremely small compression time. This
consecutive impact of the interface on the solid surface in the surface cells corresponds to the expansion
of the circular intersection between the solid surface and the surface of the liquid droplet.
The expansion speed of the circular intersection is much faster than the compression wavefront
speed, as Lesser [16] also pointed out. When the circular intersection expansion is caused purely by
compression, the circular intersection expansion speed at Dt0 after the impact VR can be written as
V(R  VDt0 )
VR ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : (4:9)
VDt0 (2R  VDt0 )

Note that s/VR is a good measure to evaluate which process dominates in the compression process, the
expansion of the circular intersection or the change in the curvature due to the compression wavefront
propagation. We evaluate s/VR, taking the limit of Dt0 to 0, obtaining
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
s V(2R  VDt0 ) s pffiffiffiffiffiffi0
lim ¼ lim Dt ¼ 0: (4:10)
Dt !0 VR
0 0
Dt !0 R  VDt0 V
We now understand that the compression deformation in the z-direction, as shown in figure 10, never
occurs because the compression wavefront should pass through a portion in the liquid droplet for that
portion to deform. We conclude that the circular intersection between the solid surface and the
surface of the liquid droplet expands at an extremely high speed, much greater than the compression
wavefront speed, compressing the spherical cap over which the circular intersection has moved while
keeping the radius of curvature of the droplet unchanged in the compression process.
Our new finding leads to the following. At the end of the droplet compression process, where neither
displacement of liquid nor propagation of compression wavelets has occurred, we assume that
the equation of the l-plane is z ¼ 0, with no change in the radius of curvature of the droplet in
the vicinity of the impact, but a compression of the spherical cap, as shown in figure 11. We set the
z-coordinate of the first impact point between the solid plane surface and the liquid droplet to be
z ¼ 2 j. We assume that no point contact can be sustained, even at the end of the compression
process; hence, the contact area between the liquid droplet and solid surface should be finite at the
end of the compression process.
We next discuss the displacement process. After the compression process, the solid plane surface, i.e.
the d-plane, moves a distance d and the l-plane moves a distance of (h 2 j ) over Dt, i.e. the l-plane
reaches z ¼ 1h at Dt, as shown in figure 11, where 1 is defined by
j
1 ¼ 1  ¼ 1  n: (4:11)
h
As we discussed above, the circular intersection expansion speed is much faster than both the
compression wavefront speed and the impact speed, especially at the very early stage of impact;
z 14

rsos.royalsocietypublishing.org
rh

l-plane

R. Soc. open sci. 5: 181101


Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

eh d-plane
h
d
r
O x

Figure 11. Cross section of a liquid droplet. The original shape of the droplet before compression is indicated by a dotted line. The
liquid cylinder with base radius rvd formed between the d- and l-planes has the same volume as that of the spherical cap whose
base is on the l-plane with height vd. We consider the limit of h to 0 as we take the limit of Dt to 0. Note that the z-direction
distance travelled by the compression wavefront generated at the compression process is uniform (the straight horizontal line on top
of the hatched region) because the circular intersection expansion speed is much faster than the speed of sound. Note that the
magnitude of j is greatly exaggerated here, i.e. j  R.

hence, the volume of the compressed spherical cap can also be much larger than the volume of the
displaced portion of the liquid droplet, which is indicated by the light grey region in figure 11. Note
that V/VR is a good measure for evaluating which process dominates the deformation of the spherical
cap, the compression process or the displacement process. We evaluate V/VR, taking the limit of Dt to
0 using equation (4.10), obtaining lim Dt!0 V/VR ¼ 0. We conclude that 1 converges to 0 in the limit
of Dt to 0, i.e. at the instant of the droplet impact, because the compression process completely
dominates the deformation of the spherical cap.
Using the notation used in §4.1, we obtain the relation vd ¼ eh; then, h is written as
v(1  aE )VDt
h¼ : (4:12)
1
0 0
We now investigate the pressure rise in the limit of Dt to 0. First, we consider the volume vm of m0 . Recall that
0 0
vm0 is the volume of a spherical cap with a height of vd and the radius of the base of rvd. vm is the volume
of a spherical segment with a radius of the base of rh and a height of h(1 2 n). Simple algebra gives
p
v0m0 ¼ r2h h(1  n2 ): (4:13)
2
0 00
Second, the volume vm of m00 is considered. Note that the upper bound of the integration in equation
(4.7) is the maximum distance travelled by the compression wave from the l-plane; hence, vbd should
be replaced by bh(1 2 n) in equation (4.7). Then, an evaluation of the integration gives
bp 2
v0m00 ¼ r h(1  n2 ): (4:14)
2 h
We estimate the average rate of change of momentum of the liquid in volume v0 , which is

(1 þ b)(1  n2 )
v0 ¼ v0m0 þ v0m00 ¼ r0 pr2h hV, (4:15)
2
to be (m0 þ m00 )aEV/Dt. Noting that instead of equation (4.1), the velocity Vl of l-plane is given by
1h eh(1  aE )
Vl ¼ ¼ V: (4:16)
Dt d
We obtain the equation for the instantaneous pressure rise DPi at the instant of impact:

r0 v0 aE V aE r0 (2  1)pr2h sVDt (2  1)
DPi ¼ lim 2
¼ lim 2 Dt
¼ lim aE r0 sV ¼ aE r0 sV: (4:17)
1!0 pr Dt
h
1!0 2 p rh
1!0 2
To obtain these results, we used the fact that 1 converges to 0 in the limit of Dt to 0. If the compression 15
process completely dominates, as in the case we examined, 1 becomes 0. On the contrary, if the

rsos.royalsocietypublishing.org
displacement process completely dominates, 1 could become 1. When 1 is 1, the coefficient of r0sV in
equation (4.17) becomes aE/2, which gives equation (1.1). We found that the coefficient 12 appears
when the displacement process completely dominates or the radial flow becomes prominent. Recall
that we found in §3.2 that a is 0.455 when the liquid in a significantly large Cell-1 can flow. We hence
conclude that the coefficient 12 that appears in equation (1.1) originates from the spherical shape of the
liquid droplet that can allow the radial flow. We emphasize that a can be 12 only when 1 ¼ 1; however,
a cannot be 12 in the limit of Dt to 0, rather a is 1, i.e. 1 is 0. Indeed, 1 can be nothing but 0 at the instant of
the droplet impact.
Note that the traversal of the compression wavefront in the droplet may be severely impulsive; hence,
the divergence of the compression wave may be negligibly small. Then, we may regard aE as 1, which

R. Soc. open sci. 5: 181101


also enforces the point impact because h in equation (4.12) converges to 0 in the limit of aE to 1. Then,
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

we finally obtain
DPi ¼ r0 sV: (4:18)

We have successfully extended the equation for the maximum pressure rise DPE developed at droplet
impact given by equation (4.8) in the limit of Dt to 0, to obtain the equation for the instantaneous
maximum pressure rise DPi at the instant of impact given by equation (4.18).
Equation (4.18) is exactly the same as equation (2.11), which predicts the impact pressure rise under
the assumption of a one-dimensional flow. Equation (4.10) shows that the central pressure develops
much more slowly than the neighbouring contact between the gas –liquid interface at the instant of
liquid droplet impact, i.e. in the limit of Dt0 to 0, as we have observed in the numerical results in §3.4;
hence, the pressure development on the z-axis is forced to be confined to one dimension, as we have
discussed in §3.3. Therefore, we conclude that the instantaneous pressure rise due to the spherical
liquid impact can be well predicted by equation (4.18).

5. Conclusion
The prime objective of this study was to answer how high is the pressure that develops at the instant of a
liquid droplet impact on a solid plate. To achieve this objective, we investigated the following issue:
At the instant of impact, is the assumption that the generated pressure coincides with the water-hammer
pressure rise relevant? The water-hammer theory and the shock relation yield the one-dimensional
liquid impact pressure rise DPs:
DPs ¼ r0 sV, (5:1)

where r0 is the density of the undisturbed liquid, s is the compression wavefront propagation speed in
the disturbed liquid and V is the impact velocity. We showed both numerically and analytically that the
liquid droplet impact pressure, which is referred to as the pressure developed at the instant of the liquid
droplet impact on the contact point, can be predicted by equation (5.1).
We examined the pressure rise at the impact of the liquid droplet by numerically solving two-
dimensional axisymmetric compressible Euler equations using the level set ghost fluid method and
the stiffened-gas equation of state. We investigated the pressure of the fluid in the surface cells that
are adjacent to the solid surface. We defined Cell-1, which is the surface cell that contains the z-axis,
and the central pressure, which is the pressure in Cell-1. We found that the maximum central pressure
rise DPmax depends on the calculation cell size, and it converges to the one-dimensional liquid impact
pressure rise DPs (equation (5.1)) as the cell size decreases. When the non-dimensional calculation cell
size Dr , which is the dimensional cell size divided by the droplet radius, is 2215, DPmax reaches
0.994DPs. We also found that the elapsed time for the pressure to reach the maximum from the instant
of impact te decreases as Dr decreases and converges to 0.
We found that the maximum central pressure DPmax is greatly affected by the status of Cell-2, which is
the surface cell adjacent to Cell-1, while the central pressure increases. When the cell size is sufficiently
small, the pressure in Cell-2 begins to rise immediately after the pressure in Cell-1 begins to rise. They rise
practically simultaneously. When the fluid in Cell-2 remains a gas after the pressure rises in Cell-1, the
liquid in Cell-1 is confined to the space bounded by the gas in Cell-2; hence, the liquid in Cell-1 can
be compressed in directions both normal and parallel to the solid surface. In other words, the liquid
in Cell-1 can flow in a direction parallel to the solid surface. When the liquid in Cell-1 cannot flow,
the flow can be assumed to be one-dimensional. Therefore, DPmax can reach DPs
We then discussed whether the impact pressure rise can be reduced from stiffened-gas pressure rise 16
DPsg when the interface that impacts on a solid surface has a finite radius of curvature by using the

rsos.royalsocietypublishing.org
numerical analysis. Numerical calculations were carried out under the conditions Dr ¼ 215 and
Dt ¼ 3:73  1010 . Under these conditions, we successfully reproduced the droplet impacts with a
solid surface with a finite radius curvature because no simultaneous interface impacts occurred in any
adjacent cells. The maximum central pressure rise DPmax was 0.994DPs. We hence concluded that the
maximum central pressure rise that develops when a droplet impacts on a solid surface can be
accurately predicted by equation (5.1), even with a finite radius of curvature interface. We emphasize that
it requires extremely high resolution in both the spatial and temporal domains to obtain the numerical
result that the maximum central pressure rise can be predicted by equation (2.10). We are convinced
that insufficient resolutions in either the numerical or experimental studies caused the previous
confusion in determining the factor.

R. Soc. open sci. 5: 181101


We shed light on the prediction of the maximum pressure rise DPE by [6],
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

aE
DPE ¼ r sV, (5:2)
2 0
where, for comparison, the compression wavefront propagation speed s is used instead of c0, the sound
velocity in the undisturbed liquid that was used in the original equation. It is widely recognized that
aE/2 comes from the spherical shape of the liquid droplet; however, our numerical analysis showed
that the impact pressure rise due to the spherical liquid droplet can be predicted by equation (5.1),
where no shape factor is required. We discussed the origin of the mysterious factor aE/2; then,
revealed the paradox.
We realized that Engel [6] derived equation (5.2) for the maximum pressure rise developed at finite time
Dt after a liquid droplet impacts on a solid surface, not for the maximum pressure rise at the instant of
impact, and that she did not take into account that the progression speed of the contact between the gas–
liquid interface and the solid surface is much faster than the compression wavefront propagation speed
at the instant of the impact. We corrected Engel’s elastic impact model taking the limit of dt to 0 by
carefully examining the compression of the cap of the spherical liquid droplet at the compression process
at the instant of impact. We found that the factor aE/2 appearing in equation (5.2) converges to 1, at the
instant of impact in the limit of impact duration time Dt to 0, because the central pressure develops much
more slowly than the neighbouring contact progresses between the gas–liquid interface and the solid
surface. We conclude that we successfully proved that the corrected Engel’s theory can also provide the
maximum central pressure rise DPs, equation (5.1), at the instant of impact between a spherical liquid
droplet which keeps its spherical shape at the instant of the impact and a rigid solid surface.
Data accessibility. Our codes, the output data and the data used to generate the figures in this work are publicly available
at https://github.com/MasaoWatanabe/Liquid-Droplet-Impact-Pressure.
Authors’ contributions. Y.T. and M.W. both designed the research methodology. Y.T. produced all figures and numerical
data. All four authors contributed to the interpretation of the results and the preparation of the manuscript.
Competing interests. We declare we have no competing interests.
Funding. The authors acknowledge the support of Japan Society for the Promotion of Science. This research is supported
by JSPS KAKENHI grant numbers JP26289031, JP16K06073 and JP17H03168.
Acknowledgements. M.W. and K.K. thank Y. Jinbo for helpful discussions.

References
1. Jenkins DC, Booker JD. 1960 The impingement 4. Rein M. 1993 Phenomena of liquid drop impact Sci. Tech. 47, 314–321. (doi:10.1080/18811248.
of water drops on a surface moving at high on solid and liquid surfaces. Fluid Dyn. Res. 12, 2010.9711960)
speed. In Symp. Aerodynamic Capture of 61– 93. (doi:10.1016/0169-5983(93)90106-K) 8. Yarin AL. 2006 Drop impact dynamics:
Particles (ed. EG Richardson), pp. 97 –103. 5. Field JE, Camus JJ, Tinguely M, Obreschkow D, splashing, spreading, receding, bouncing.
Leatherhead, UK: British Coal Utilisation Farhat M. 2012 Cavitation in impacted drops Ann. Rev. Fluid Mech. 38, 159 –192. (doi:0.
Research Association. and jets and the effect on erosion damage 1146/annurev.fluid.38.050304.092144)
2. Bowden FP, Field JE. 1964 The brittle fracture of thresholds. Wear 290 –291, 154– 160. (doi:10. 9. Obreschkow D, Dorsaz N, Kobel P, De Bosset A,
solids by liquid impact, by solid impact, and by 1016/j.wear.2012.03.006) Tinguely M, Field JE, Farhat M. 2011 Confined
shock. Proc. R. Soc. Lond. A 282, 331–352. 6. Engel OG. 1955 Waterdrop collision with shocks inside isolated liquid volumes: a new
(doi:10.1098/rspa.1964.0236) solid surfaces. J. Res. Natl Bur. Stand. 55, path of erosion? Phys. Fluids 23, 101702.
3. Brunton JH, Camus JJ. 1970 The flow of a liquid drop 281–298. (doi:10.1063/1.3647583)
during impact. In Proc. 3rd Int. Conf. Rain Erosion and 7. Xiong J, Koshizuka S, Sakai M. 2010 Numerical 10. Engel OG. 1960 Note on particle velocity in collisions
Associated Phenomena (ed. CM Preece), pp. 327– analysis of droplet impingement using the between liquid drops and solids. J. Res. Natl Bur.
352. Farnborough, UK: Royal Aircraft Establishment. moving particle semi-implicit method. J. Nucl. Stand. 64, 497–498. (doi:10.6028/jres.064a.048)
11. Rochester MC. 1977 The impact of a liquid drop impact. Ocean Eng. 72, 98 –106. (doi:10.1016/ Information Service, US Department of 17
with a solid surface and the effects of the j.oceaneng.2013.06.012) Commerce.
properties of the liquid on the erosion of solids. 29. Cooker JC, Peregrine DH. 1995 Pressure-impulse 46. Saurel R, Abgrall R. 1999 A simple method for

rsos.royalsocietypublishing.org
PhD thesis, University of Cambridge, UK. theory for liquid impact problems. J. Fluid. compressible multifluid flows. SIAM J. Sci.
12. Rochester M, Brunton J. 1979 Pressure Mech. 297, 193–214. (doi:10.1017/ Comput. 21, 1115 –1145. (doi:10.1137/
distribution during drop impact. In Proc. 5th Int. S0022112095003053) S1064827597323749)
Conf. on Erosion by Liquid and Solid Impact (ed. 30. Philippi J, Lagrée PY, Antkowiak A. 2015 A 47. Cocchi JP, Saurel R. 1997 A Riemann problem
JE Field), pp. 6-1– 6-7. Cambridge, UK: pressure impulse theory for hemispherical liquid based method for the resolution of
Cavendish Laboratory. impact problems. Eur. J. Fluid. Mech. B/Fluids compressible multimaterial flows. J. Comp.
13. Bourne NK. 2005 On impacting liquid jets and 67, 417– 426. (doi:10.1017/jfm.2016.142) Phys. 137, 265–298. (doi:10.1006/jcph.1997.
drops onto polymethylmethacrylate targets. 31. Wagner H. 1932 Über Stoß-und Gleitvorgänge 5768)
Proc. R. Soc. A 461, 1129 – 1145. (doi:10.1098/ an der Overfläche von Flüssigkeiten. Z. Angew. 48. Shyue KM. 1999 A fluid-mixture type algorithm
rspa.2004.1440) Math. Mech. 12, 193–215. (Trans. Phenomena for compressible multicomponent flow with
14. Heymann FJ. 1969 High-speed impact between associated with impacts and sliding on liquid Mie-Güneisen equation of state. J. Comp. Phys.

R. Soc. open sci. 5: 181101


a liquid drop and a solid surface. J. Appl. Phys. surfaces, NACA Trans. 1366). (doi:10.1002/zamm. 156, 43 –88. (doi:10.1006/jcph.2001.6801)
40, 5113 –5122. (doi:10.1063/1.1657361) 19320120402) 49. Johnsen E. 2007 Numerical simulations of non-
Downloaded from https://royalsocietypublishing.org/ on 27 December 2021

15. Heymann FJ. 1968 On the shock wave velocity 32. Howison SD, Ockendon JR, Wilson SK. 1991 spherical bubble collapse. PhD thesis, California
and impact pressure in high-speed liquid-solid Incompressible water-entry problems at small Institute of Technology.
impact. J. Basic Eng-T ASME 90, 400– 402. deadrise angles. J. Fluid Mech. 222, 215 –230. 50. Tatekura Y, Fujikawa T, Jinbo Y, Sanada T,
(doi:10.1115/1.3605114) (doi:10.1017/S0022112091001076) Kobayashi K, Watanabe M. 2015 Observation of
16. Lesser MB. 1981 Analytic solutions of 33. Korobkin AA, Scolan YM. 2001 Three- water-droplet impacts with velocities of
liquid-drop impact problems. Proc. R. Soc. dimensional theory of water impact. Part O(10 m s21) and subsequent flow field. ECS
Lond. A 377, 289–308. (doi:10.1098/rspa. 1. Inverse Wagner problem. J. Fluid Mech. 440, J. Solid State Sci. Technol. 4, N117 –N123.
1981.0125) 293–326. (doi:10.1017/S002211200100475X) (doi:10.1149/2.0111509jss)
17. Field JE. 1999 ELSI conference invited lecture: 34. Wilson SK. 1991 A mathematical model for the 51. Sussman M, Smereka P, Osher S. 1994 A level
liquid impact: theory, experiment, applications. initial stages of fluid impact in the presence of a set approach for computing solutions to
Wear 233 – 235, 1 –12. (doi:10.1016/S0043- cushioning fluid layer. J. Eng. Math. 25, incompressible two-phase flow. J. Comp. Phys.
1648(99)00189-1) 265–285. (doi:10.1007/BF00044334) 114, 146– 159. (doi:10.1006/jcph.1994.1155)
18. Korobkin AA. 1992 Blunt-body impact on a 35. Hicks PD, Purvis R. 2011 Air cushioning in 52. Sethian JA. 1996 A fast marching level set
compressible liquid surface. J. Fluid Mech. 244, droplet impacts with liquid layers and other method for monotonically advancing fronts.
437–453. (doi:10.1017/S0022112092003136) droplets. Phys. Fluids 23, 062104. (doi:10.1063/ Proc. Natl Acad. Sci. USA 93, 1591 –1595.
19. Korobkin AA. 1997 Asymptotic theory of liquid- 1.3602505) (doi:10.1073/pnas.93.4.1591)
solid impact. Phil. Trans. R. Soc. Lond. A 355, 36. van der Veen RCA, Tran T, Lohse D, Sun C. 2012 53. Osher S, Fedkiw RP. 2001 Level set methods:
507–522. (doi:10.1098/rsta.1997.0021) Direct measurements of air layer profiles under an overview and some recent results. J. Comp.
20. Hwang JBG, Hammitt FG. 1977 High-speed impacting droplets using high-speed color Phys. 161, 463–502. (doi:10.1006/jcph.
impact between curved liquid surface and rigid interferometry. Phys Rev. E 85, 026315. (doi:10. 2000.6636)
flat surface. J. Fluids Eng. 99, 396– 404. (doi:10. 1103/PhysRevE.85.026315) 54. Fedkiw RP. 2002 Coupling an Eulerian fluid
1115/1.3448774) 37. Li EQ, Vakarelski IU, Thoroddsen ST. 2015 calculation to a Lagrangian solid calculation
21. Haller KK. 2002 High-velocity impact of a liquid Probing the nanoscale: the first contact of an with the ghost fluid method. J. Comp. Phys.
droplet on a rigid surface: the effect of liquid impacting drop. J. Fluid Mech. 785, R2. (doi:10. 175, 200– 224. (doi:10.1006/jcph.2001.6935)
compressibility. PhD thesis, Swiss Federal 1017/jfm.2015.643) 55. Harten A, Engquist B, Osher S, Chakravarthy SR.
Institute of Technology, Zurich. 38. Josserand C, Thoroddsen ST. 2016 Drop impact 1987 Uniformly high order accurate schemes,
22. Haller KK, Ventikos Y, Poulikakos D, Monkewitz on a solid surface. Ann. Rev. Fluid Mech. 48, essentially non-oscillatory schemes, III. J. Comp.
P. 2002 Computational study of high-speed 365–391. (doi:10.1146/annurev-fluid-122414- Phys. 71, 231–303. (doi:10.1016/0021-
liquid droplet impact. J. Appl. Phys. 92, 034401) 9991(87)90031-3)
2821 –2828. (doi:10.1063/1.1495533) 39. Cook SS. 1928 Erosion by water-hammer. 56. Shu CW, Osher S. 1989 Efficient implementation
23. Sanada T, Watanabe M, Shirota M, Yamase M, Proc. R. Soc. Lond. A Math. Phys. Sci. 119, of essentially nonoscillatory schemes, II.
Saito T. 2008 Impact of high-speed steam- 481–488. (doi:10.1098/rspa.1928.0107) J. Comp. Phys. 83, 32– 78. (doi:10.1016/0021-
droplet spray on solid surface. Fluid Dyn. Res. 40, 40. Gardner FW. 1932 The erosion of steam turbine 9991(89)90222-2)
627–636. (doi:10.1016/j.fluiddyn.2007.12.014) blades. The Engineer 153, 146 –147, 74– 176, 57. Jiang GS, Peng D. 2000 Weighted ENO schemes
24. Fedkiw RP, Aslam T, Merriman B, Osher S. 1999 202–205. for Hamilton –Jacobi equations. SIAM J. Sci.
A nonoscillatory Eulerian approach to interfaces 41. de Haller P. 1933 Untersuchungen über die Comput. 21, 2126 –2143. (doi:10.1137/
in multimaterial flows (the ghost fluid method). durch kavitation hervorgerufenen korrosionen. S106482759732455X)
J. Comp. Phys. 152, 457– 492. (doi:10.1006/ Schweiz. Bauztg. 101, 243– 246, 260–264. 58. Okorn-Schmidt HF et al. 2014 Particle cleaning
jcph.1999.6236) (doi:10.5169/seals-82996) technologies to meet advanced semiconductor
25. Sanada T, Ando K, Colonius T. 2011 A 42. Brunton JH, Rochester MC. 1979 Erosion of solid device process requirements. ECS J. Solid State
computational study of high speed droplet surfaces by the impact of liquid drops. In Sci. Technol. 3, N3069 –N3080. (doi:10.1149/
impact. Fluid Dyn. Mater. Process. 7, 329 –340. Erosion: treatise on materials science and 2.011401jss)
(doi:10.3970/fdmp.2011.007.329) technology, vol. 16 (ed. CM Preece), 59. Xu K et al. 2009 Removal of nano-particles by
26. Kondo T, Ando K. 2016 One-way-coupling pp. 185–248. London, UK: Academic Press. aerosol spray: effect of droplet size and velocity
simulation of cavitation accompanied by high- 43. Lesser MB, Field JE. 1983 The impact of on cleaning performance. Solid State Phenom.
speed droplet impact. Phys. Fluids 28, 033303. compressible liquids. Ann. Rev. Fluid Mech. 15, 145 –146, 31 –34. (doi:10.4028/www.
(doi:10.1063/1.4942894) 97– 122. (doi:10.1146/annurev.fl.15.010183. scientific.net/SSP.145-146.31)
27. Johnsen E, Colonius T. 2006 Implementation of 000525) 60. Camus JJ. 1971 A study of high-speed liquid
WENO schemes in compressible 44. Engel OG. 1973 Damage produced by high- flow in impact and its effect on solid surfaces.
multicomponent flow problems. J. Comp. Phys. speed liquid-drop impacts. J. Appl. Phys. 44, PhD thesis, University of Cambridge, UK.
219, 715– 732. (doi:10.1016/j.jcp.2006.04.018) 692–704. (doi:10.1063/1.1662246) 61. Field JE, Dear JP, Ogren JE. 1989 The effects of
28. Hsu CY, Liang CC, Teng TL, Nguyen AT. 2013 A 45. Harlow FH, Amsden AA. 1971 Fluid dynamics: a target compliance on liquid drop impact. J. Appl.
numerical study on high-speed water jet LASL monograph LA-4700. National Technical Phys. 65, 533–540. (doi:10.1063/1.343136)

You might also like