ArticlePDF Available

The rigorous derivation of Young, Cassie–Baxter and Wenzel equations and the analysis of the contact angle hysteresis phenomenon

Authors:

Abstract and Figures

The rigorous derivation of Young, Cassie–Baxter and Wenzel equations carried out in the framework of the unified thermodynamic approach is presented. Wetting of rough surfaces controlled with external stimuli is treated. Areas of validity of Cassie–Baxter and Wenzel approaches are discussed. General properties of the contact angle hysteresis are investigated on the same thermodynamic basis.
Content may be subject to copyright.
This article was published in an Elsevier journal. The attached copy
is furnished to the author for non-commercial research and
education use, including for instruction at the author’s institution,
sharing with colleagues and providing to institution administration.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy
The rigorous derivation of Young, Cassie–Baxter and Wenzel
equations and the analysis of the contact angle hysteresis phenomenon
Gene Whyman, Edward Bormashenko
*
, Tamir Stein
The Ariel University Centre of Samaria, P.O. Box 3, Ariel 440700, Israel
Received 28 August 2007; in final form 13 November 2007
Available online 19 November 2007
Abstract
The rigorous derivation of Young, Cassie–Baxter and Wenzel equations carried out in the framework of the unified thermodynamic
approach is presented. Wetting of rough surfaces controlled with external stimuli is treated. Areas of validity of Cassie–Baxter and Wen-
zel approaches are discussed. General properties of the contact angle hysteresis are investigated on the same thermodynamic basis.
2007 Elsevier B.V. All rights reserved.
1. Introduction
Wetting phenomena are essential and ubiquitous in a
variety of natural and technological processes. Wetting of
solids has been exerted to intensive experimental and theo-
retical activities in two past decades [1–4]. An interest to
the wetting has been also increased due to the superhydro-
phobicity phenomenon, revealed recently in a variety of
natural and artificial objects [5–10].
Wetting of smooth and rough solids is governed by
Young, Wenzel and Cassie–Baxter equations [1,2,10–13].
The Young equation results from the equilibrium of forces
acting on the triple line [1]. Wenzel and Cassie–Baxter
equations supplying the value of apparent contact angles
inherent to rough surfaces for a long time were based on
the semi-empirical arguments. The rigorous derivation of
these equations has been carried out recently basing on
the principle of virtual work and concepts supplied by non-
extensive thermodynamics [14,15]. The factor complicating
the understanding of wetting phe nomena is the effect of the
hysteresis of the contact angle; actually at the same (even
smooth) surface co-exist various contact angles, the mini-
mal and maximal of which are called receding and advan c-
ing contact angles respectively [16–24].
The profound microscopic understanding of the phe-
nomenon of the contact angle hysteresis could be picked
up from the recent work by Yaminsky [18]. According to
Yaminsky the phenomenon of the contact angle hysteresis
resembles the well-known effect of the static friction, when
the triple line is pinned to the solid substrate due to the
long-range forces or roughness of the surface [18]. Thus
the spectrum of contact angles becomes possibl e [1,16–24].
However the drop when deposited on the surface dem-
onstrates usually contact angles dictated by Young, Wenzel
or Cassie–Bax ter equations; thus the open question is: why
the drop ‘selects’ certain contact angle from the spectrum
of possible ones. We demonstrate in our communication
that contact angles derived from Young, Wenzel or
Cassie–Baxter equations supply minimum to the free
energy of the drop. Thus one more rigorous derivation of
these equations based on pure thermodynamic arguments
becomes possible.
2. Derivation of the Young formula
Let us start with the drop of the radius R deposited on
the ideally flat surface. If h denotes contact angle we have
the volume V and surface S of the liquid–air interface
expressed as (Fig. 1)
V ¼
pR
3
3
ð1 cos hÞ
2
ð2 þ cos hÞ; ð1Þ
0009-2614/$ - see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.cplett.2007.11.033
*
Corresponding author. Fax: +972 3 9066621.
E-mail address: Edward@ariel.ac.il (E. Bormashenko).
www.elsevier.com/locate/cplett
Available online at www.sciencedirect.com
Chemical Physics Letters 450 (2008) 355–359
Author's personal copy
S ¼ 2pR
2
ð1 cos hÞ: ð2Þ
The Gibbs free energy of the drop is expressed by Eq. (3)
(within an additive constant)
G ¼ cS þ pðR sin hÞ
2
ðc
SL
c
SV
Þ¼cS p ðR sin hÞ
2
a; ð3Þ
where c, c
SL
, c
SV
are the surface tensions at the liquid/
vapor, solid/liquid and solid/vapor interfaces, respectively,
S is the spherical liquid/air interface area. Having in mind
further applications we have introduced here the constant
a, which in the special case of flat homogeneous substrates
is defined as
a ¼ c
SV
c
SL
: ð4Þ
The general form of the dependence in the right-hand part
of Eq. (3) remains true in many other cases of physical
interest.
Now we will make the main assumption of our treat-
ment, namely we assume the constant volume of the drop ,
V = constant. Substitution of formulae (1) and (2) into Eq.
(3) yields
G ¼
9pV
2
ð1 cos hÞð2 þ cos hÞ
2
"#
1=3
ð2c að1 þ cos hÞÞ: ð5Þ
Now G is a function of only one independent variable h
that is the contact angle.
The straightforward differentiation gives
dG
dh
¼
9V
2
p
ð1 cos hÞ
4
ð2 þ cos hÞ
5
"#
1
3
2ða c cos hÞ sin h: ð6Þ
It is clear that
dG
dh
ðh ¼ h
E
Þ¼0 is fulfilled when
a ¼ c cos h
E
; or ð7Þ
cos h
E
¼
c
SV
c
SL
c
: ð8Þ
Thus the well-known Young equation for the equilibrium
contact angle on flat homogeneous surfaces is obtained.
3. The Cassie–Baxter and Wenzel equations
Now consider, following Wenzel [1] , the rough surface
(Fig. 1B) characterized by the roughness f > 1, i.e. by the
ratio of the real surface in contact with liquid to its projec-
tion onto the horizontal plane. This means that the area of
liquid–solid interface is equal to p(Rsin h)
2
f and we have for
the free energy equation (3) with
a ¼ f ðc
SV
c
SL
Þ: ð9Þ
From Eq. (7) then follows the Wenzel equation for the
equilibrium contact angle h
*
called the apparent contact
angle
cos h
¼
a
c
¼ f
ðc
SV
c
SL
Þ
c
¼ f cos h
E
: ð10Þ
Let suppose that the surface under the drop is flat but con-
sists of n sorts of materials randomly distributed over the
substrate. Each material is characterized by its own surface
tension coefficients c
i, SL
and c
i, SV
, and by the fraction U
i
in
the substrate surface, U
1
+ U
2
+ + U
n
= 1. Analogously
to the above treatment we sim ply put in Eq. (3)
a ¼
X
n
1
U
i
ðc
i;SV
c
i;SL
Þð11Þ
and obtain the Cassie–Baxter apparent contact angle
c cos h
¼
X
n
1
U
i
ðc
i;SV
c
i;SL
Þ: ð12Þ
The most frequently n =2
c cos h
¼ U
1
ðc
1;SV
c
1;SL
Þþð1 U
1
Þðc
2;SV
c
2;SL
Þ: ð13Þ
A more special case is important when air remains trapped
under the drop (Fig. 1C). If the air surface fraction under
the drop is 1 U, then c
2,SV
= 0 and c
2,SL
= c (S V)
a ¼ c cos h
¼ Uðc
SV
c
SL
Þð1 UÞc: ð14Þ
Let find the equilibrium surface energy of the drop on the
substrate. Substituting Eq. (7) in the form a = c cosh
*
into
Eq. (5) one obtains
G
0
¼ c½ð3V Þ
2
pð1 cos h
Þ
2
ð2 þ cos h
Þ
1=3
ð15Þ
The graph of G
0
/c as a function of cos h
*
= a/c is presented
in Fig. 2 . The value of G
0
at h
*
= p is equal to ð36V
2
pÞ
1=3
c ¼ 4pR
2
0
c that is the surface energy of the spherical drop
of the radius R
0
which is in contact with vapor only (before
depositing on a substrate). The parameter a is defined by
the geometry of the substrate and by the surface tension
coefficients. Since G
0
is a decreasing function of a, a system
with flexible parameters will tend to increase a. Practically
the wetting flexibility of the system could be achieved with
external stimuli, such as temperature, external fields or
vibration of the drop; this was already realized in the elec-
trowetting or the recently reported vibration experiments
[25,26]. The additional restriction is given by Eq. (12):
Fig. 1. Schemes of different wetting regimes. A flat substrate; B rough substrate, the Wenzel regime; C rough substrate with air trapped under the
drop, the Cassie–Baxter regime.
356 G. Whyman et al. / Chemical Physics Letters 450 (2008) 355–359
Author's personal copy
a 6 c. For example, if the fraction U of wetted surface of
the system described by Eq. (13) may change, the system
will proceed to the state with a = min(c, c
SV
c
SL
) with
the maximal possible U supplying maximum to a in Eq.
(14) (usually c
SV
c
SL
> 0). If c > c
SV
c
SL
then
a = c
SV
c
SL
, U = 1 that means that air is completely re-
moved from under the drop and the contact angle receives
its (finite) Young value. In the opposite case c < c
SV
c
SL
,
a = c, h
*
= 0 and the substrate is completely wetted by the
liquid but air pockets of total area 1 U remain trapped
under the liquid film:
1 U ¼
c
SV
c
SL
c
c
SV
c
SL
þ c
: ð16Þ
The similar techni que based on Eqs. (9)–(14) may be ap-
plied for the analysis of wetting the surfaces of a compli-
cated structure, superhydrophobicity, etc.
It has to be emphasized that the derivation of the Cassie–
Baxter and Wenzel eq uations presented in our communica-
tion suffers from the same shortc oming like those carried
out by Bico et al. based on the virtual work principle
[14]. Both our approach and application of the virtual
work principle presumes the free displacement of the triple
line, however in the situation depicted in Fig. 3 when water
fills the grooves of the surface relief, state A is statically
impossible. The same is true for the situation of heteroge-
neous wetting, when air pock ets are formed (see Fig. 3D
and E). In this case the equilib rium of the triple line
becomes possible only when the drop sits on solid islands
only (see Fig. 3F). The equilibrium in states D and E is
impossible. The drop can sit on the air pocket but the triple
line cannot. Thus free displacement of the triple line
becomes impossible and both application the virtual work
principle and variation of contact angle are at least prob-
lematic. In this situation the detailed study of the fine struc-
ture of triple line, taking into account formation of
precursor film surrounding the drop is necessary [27,28].
On the other hand the approach presented in our com-
munication explains why the drop ‘selects certain contact
angle from the variety of angles, permitted by the contact
angle hysteresis phenomenon. The equilibrium contact
angle supplies minimum value to the free energy of the
drop.
4. Contact angle hysteresis
Actually, the experimentally established apparent contact
angle h
*
is confined within a certain range h
rec
< h
*
< h
adv
,
where h
adv
and h
rec
are the so-called advancing and reced-
ing contact angles [1,16–21]. The difference between h
adv
and h
rec
is called the contact angle hysteresis (CAH, see
Section 1). Some general properties of this phenomenon
may be understood on the basis of the above used simple
thermodynamic model.
The straightforward differentiation of Eq. (6) immedi-
ately yields
d
2
G
dh
2
h¼h
E
¼
9V
2
p
ð1 cos h Þ
4
ð2 þ cos h Þ
5
"#
1
3
2c sin
2
h: ð17Þ
The increase in the energy in Eq. (5) is for small variations
dh of the contact angle
d
2
E ¼
1
2
d
2
G
dh
2
h¼h
E
dh
2
: ð18Þ
Now we apply the approach proposed by Shanahan for
consideration of water drop evaporation [22]. It has to be
emphasized that the CAH has been observed experimen-
tally in two different experimental situations, namely, when
the contact line is pinned [18,23,24] and when contact line
moves [22,26]. We will restrict our treatment with the CAH
due to displacement of triple line. We also suppose that this
displacement is slow, thus effects due to ‘dynamic contact
angle phenomena’ are neglected [1]. Let U be the height
0
0.5
1
1.5
2
2.5
-1.0 -0.5 0.0 0.5 1.0
cosθ*
G
0
/γ (3V)
2/3
0
0.5
1
1.5
2
2.5
h (θ*)
Fig. 2. The equilibrium surface energy of the sessile drop (solid line) and
geometrical factor h(h) of the contact angle hysteresis (dashed line) versus
the apparent contact angle.
Fig. 3. Various situations of wetting for the Wenzel (A,B,C) and Cassie–Baxter (D, E, F) states.
G. Whyman et al. / Chemical Physics Letters 450 (2008) 355–359 357
Author's personal copy
of the potential barrier connected with the motion of the
triple line along a substrate. This quantity should obviously
be related to the unit length of this line. Thus the move-
ment will start after the energy increase reaches the poten-
tial barrier U
d
2
E
2pR sin h
¼ U : ð19Þ
The natural estimation of the CAH is h
adv
h
rec
=2dh,
thus from this and Eqs. (17)–(19), (1), the final result
follows:
h
adv
h
rec
¼
8U
cR
0

1=2
hðh
Þ;
hðh
Þ¼
ð1 cos h
Þ
1=12
ð2 þ cos h
Þ
2=3
2
1=3
ð1 þ cos h
Þ
1=4
; ð20Þ
where R
0
=(3V/4p)
1/3
is the initial radius of the spherical
drop before deposition on the substrate. The function
h(h) of the equilibrium contact angle is shown in Fig. 2.
Some important qualitative conclusions follow from this
result. First of all, the CAH does not depend explicitly on
the constant a, i.e. wetting constants and roughness. The
larger CAH in the Wenzel state compared to the Cassie
one is connected with the larger value of U in the former
case [29]. It should be emphasized that the result in Eq.
(20) follows from the geometric consideration and physical
information is mostly concentrated in the magnitude of the
potential barrier U. In turn, the value of U is determined by
both adhesion of liquid by solid and roughness of soli d
[20].
In the broad region of values, approximately from 50 to
140, the CAH does not depend on the equilibrium contact
angle h
*
(see Fig. 2, h(h) 1.3). But it becomes more
pronounced for angles close to 180 (so-called superhydro-
phobicity situation). Ultra-low hysteresis reported recently
for superhydrophobic surfaces is due to specific effects con-
nected with hierarchical roughness of the relief that may
also be explained by small U values [7,10].
For low contact angles Eq. (20) predicts very low CAH
(Fig. 2). Probably this is the reason for the lack of CAH
experimental data for water on metals. Fig. 2 also shows
that both the surface energy and the potential barrier tend
to zero for h
*
! 0.
The CAH dependence on the drop volume is very weak,
however for small drops the CAH h
adv
h
rec
becomes lar-
ger than h
*
itself. This supplies one of the possible explana-
tions why in the end of drop evaporation the contact angle
approaches to zero [22]. Another consequence of Eq. (20) is
that the CAH of liquids with a low surface tension should
be larger when compared to those with a high one, all other
conditions being equal, i.e. U U(c), h(h) @ constant.
The dimension of U is noteworthy. It is J/m and coin-
cides with the dimension of linear surface tension debated
intensively in the past decade [1,30,31]. Pompe and Her-
minghaus established U as 10
10
–10
11
J/m for water drop-
lets deposited on the quartz glass substrate [30]. These
values yield according to formula (20) very small CAH of
few tenths of a degree. This pr ediction coincides with
experimental findings for small CAH values observed on
the atomically flat silica substrates [32,33]. However, the
experimental data concerning CAH values inherent for
water/flat silica system are contradictory, and CAH as
much as 10 [32] and even 20 have been also reported
[34]. Large values of U corresponding to mentioned above
CAH values U 10
7
–10
6
J/m are perhaps due to con-
taminants, surface roughness or dynamic contact angle
effects. It as also noteworthy that straightforward applica-
tion of formula (20) for calculation of U needs some pre-
cautions, since it neglects various factors stipulating
CAH, including precursor film surrounding the film, heter-
ogeneity of the surface, etc.
Recently, the alternative calculation of the actuat ing
force of the drop motion on super hydrophobic textured
surfaces was done [35], which gave for U the values in
the range of 10
7
–10
5
J/m for CAH of 10–20 in agree-
ment with the result of Eq. (20). Moreover, the depen dence
of this force on the equilibrium contact angle [35] accords
with that of Eq. (20). In particular, the potential barrier
turns to zero for h
*
! 180 as U 1/h
2
(h
*
), h(h
*
) !1
(Fig. 2).
5. Conclusions
Several wetting phenomena have been considered on the
single mathematical basis. The expression for the free
energy of a sessile drop on the solid substrate was used
for derivation of the Young, Wenzel, and Cassie–Baxter
equations describing different wetting processes. Advanta-
ges and shortcomings of the existing wetting models have
been discussed. General consequences are deduced for the
contact angle hysteresis, which grows with the decrease of
the surface tension and drop volume.
References
[1] P.G. de Gennes, F. Brochard-Wyart, D. Qu
_
er
_
e, Capillarity and
Wetting Phenomena, Springer, Berlin, 2003.
[2] H.Y. Erbil, Surface Chemistry of Solid and Liquid Interfaces,
Blackwell Publishing, Oxford, 2006.
[3] P.G. de Gennes, Rev. Mod. Phys. 57 (3) (1985) 827.
[4] D. Bonn, D. Ross, Rep. Prog. Phys. 64 (2001) 1085.
[5] W. Barthlott, C. Neinhuis, Planta 202 (1997) 1.
[6] A. Shibuichi, T. Onda, N. Satoh, K. Tsujii, J. Phys. Chem. 100 (1996)
19512.
[7] S. Herminghaus, Europhys. Lett. 52 (2000) 165.
[8] E. Bormashenko, T. Stein, G. Whyman, Y. Bormashenko, R. Pogreb,
Langmuir 22 (24) (2006) 9982.
[9] E. Bormashenko, Y. Bormashenko, T. Stein, G. Whyman, E.
Bormashenko, J. Colloid Interf. Sci. 311 (1) (2007) 212.
[10] N.A. Patankar, Langmuir 20 (2004) 7097.
[11] A.B.D. Cassie, S. Baxter, Trans. Faraday Soc. 40 (1944) 546.
[12] A.B.D. Cassie, Discuss. Faraday Soc. 3 (1948) 11.
[13] L. Gao, Th. J. McCarthy, Langmuir 23 (2007) 3762.
[14] J. Bico, U. Thiele, D. Que
´
re
´
, Colloids Surf. A 206 (2002) 41.
358 G. Whyman et al. / Chemical Physics Letters 450 (2008) 355–359
Author's personal copy
[15] P. Letellier, Al. Mayaffre, M. Turmine, J. Colloid Interf. Sci. 314
(2007) 604.
[16] R.E. Johnson, R.H. Dettre, J. Chem. Phys. 68 (7) (1964) 1744.
[17] R.E. Johnson, R.H. Dettre, J. Chem. Phys. 69 (5) (1965) 1507.
[18] V.V. Yaminsky, Molecular mechanisms of hydrophobic transitions,
in: J. Drelich, J.S. Laskowski, K.L. Mittal (Eds.), Apparent and
Microscopic Contact Angles, VSP BV, Utrecht, 2000, p. 47.
[19] J. Drelich, Instability of the three-phase contact region and its effect
on contact angle relaxation, in: J. Drelich, J.S. Laskowski, K.L.
Mittal (Eds.), Apparent and Microscopic Contact Angles, VSP BV,
Utrecht, 2000, p. 27.
[20] M. Nosonovsky, J. Chem. Phys. 126 (2007) 224701.
[21] A. Marmur, Soft Matter 2 (2006) 12.
[22] M.E.R. Shanahan, Langmuir 11 (1995) 1041.
[23] C.W. Extrand, J. Colloid Interf. Sci. 207 (1998) 11.
[24] C.W. Extrand, Langmuir 20 (2004) 4017.
[25] T.N. Krupenkin, J.A. Taylor, E.N. Wang, P. Kolodner, M. Hodes,
T.R. Salamon, Langmuir 23 (18) (2007) 9128.
[26] E. Bormashenko, R. Pogreb, G. Whyman, Y. Bormashenko, M.
Erlich, Appl. Phys. Lett. 90 (2007) 201917.
[27] E. Bormashenko, Y. Bormashenko, T. Stein, G. Whyman, R. Pogreb,
Langmuir 23 (2007) 4378.
[28] M. Nosonovsky, Langmuir 23 (2007) 9919.
[29] L. Gao, Th. J. McCarthy, Langmuir 22 (2006) 2966.
[30] T. Pompe, S. Herminghaus, Phys. Rev. Lett. 85 (9) (2000) 1930.
[31] P. Blecua, R. Lipowsky, J. Kierfeld, Langmuir 22 (2006) 11041.
[32] X. Xie, N.R. Morrow, Contact angles on quartz induced by
adsorption of heteropolar hydrocarbons, in: J. Drelich, J.S. Laskow-
ski, K.L. Mittal (Eds.), Apparent and Microscopic Contact Angles,
VSP BV, Utrecht, 2000, p. 129.
[33] St. Ecke, M. Preuss, H.-J. Butt, Micro-sphere tensiometry to measure
advancing and receding contact angles on individual particles, in: J.
Drelich, J.S. Laskowski, K.L. Mittal (Eds.), Apparent and Micro-
scopic Contact Angles, VSP BV, Utrecht, 2000, p. 377.
[34] L.C. Salay, A.M. Carmona-Ribeiro, Wetting of SiO
2
surfaces by
phospholipids dispersions, in: J. Drelich, J.S. Laskowski, K.L. Mittal
(Eds.), Apparent and Microscopic Contact Angles, VSP BV, Utrecht,
2000, p. 95.
[35] J. Yang, J.C. Chen, K.-J. Huang, J.A. Yeh, J. Microelectromech.
Syst. 15 (2006) 697.
G. Whyman et al. / Chemical Physics Letters 450 (2008) 355–359 359
... These results align with the conclusions of studies by researchers such as Ridgway et al. and other works referenced in [47][48][49][50][51][52], which demonstrated that modifications in membrane surface roughness can significantly impact fouling behavior and surface energy calculations. Ridgway's study on bacterial adhesion highlighted the critical role of surface Figure 9 shows the cosine of the contact angle plotted against the surface tension for the PVDF-HFP and PVDF-HFP/Zn(4-PPOx) 4 Pc membranes. ...
... These results align with the conclusions of studies by researchers such as Ridgway et al. and other works referenced in [47][48][49][50][51][52], which demonstrated that modifications in membrane surface roughness can significantly impact fouling behavior and surface energy calculations. Ridgway's study on bacterial adhesion highlighted the critical role of surface roughness in biofilm formation, emphasizing that hydrophobic interactions between bacterial cell surface components and the membrane surface are crucial during the initial stages of adhesion and biofilm development. ...
Article
Full-text available
Fouling and biofouling remain significant challenges in seawater desalination plants. One practical approach to address these issues is to develop anti-biofouling membranes. Therefore, novel hybrid zinc phthalocyanine/polyvinylidene fluoride-co-hexafluoropropylene (Zn(4-PPOx)4Pc/PVDF-HFP) membranes were prepared by electrospinning to evaluate their properties against biofouling. The hybrid nanofiber membrane was characterized by atomic force microscopy (AFM), attenuated total reflectance-Fourier transform infrared (ATR-FTIR) spectroscopy, and contact angle measurements. The theoretical calculations of PVDF-HFP, Zn(4-PPOx)4Pc), and Zn(4-PPOx)4Pc/PVDF-HFP nanofibers were performed using a hybrid functional RB3LYP and the 6-31 G (d,p) basis set, employing Gaussian 09. DFT calculations illustrated that the calculated physical and electronic parameters ensured the feasibility of the interaction of PVDF-HFP with Zn(4-PPOx)4Pc via a halogen–hydrogen bond, resulting in a highly stable and remarkably reactive structure. Moreover, molecular electrostatic potential (MEP) maps were drawn to identify the reactive regions of the Zn(4-PPOx)4Pc and PVDF-HFP/Zn(4-PPOx)4Pc nanofibers. Molecular docking analysis revealed that Zn(4-PPOx)4Pc has highest binding affinity (−8.56 kcal/mol) with protein from S. aureus (1N67) mainly with ten amino acids (ASP405, LYS374, GLU446, ASN406, ALA441, TYR372, LYS371, TYR448, LYS374, and ALA442). These findings highlight the promising potential of Zn(4-PPOx) 4Pc/PVDF-HFP nanocomposite membranes in improving the efficiency of water desalination by reducing biofouling and providing antibacterial properties.
... In 1805 Young found the relationship between the contact angle and the surface free energy of the solid. Young's studies form the basis of the studies on the wettability behavior of materials [20]. The contact angle is the angle formed by the materials in the solid, liquid and gas phases together with the effect of gravity. ...
Article
Full-text available
The solid material's wettability behavior is determined by the contact angle value. The wettability of materials is critical in the coating, lubricating, and insulating industries. In high-tech industries including corrosion resistance, water-oil separation, medical material production, implant technology, and friction reduction, the contact angle is very important. A new method called the full angle method was used to measure the contact angle and compared with current methods. The wettability behavior of plexiglass, which is employed in a variety of applications ranging from lighting to decorating, and industrial designs to accessory production, was explored in this study. The measurement of the contact angle was done by dropping 1.8 M (molar) of saltwater over the plexiglass materials. In order to examine the effect of evaporation on the contact angle, changes in contact angle, height and baseline were investigated depending on the waiting time.
... e., the complete wetting state) is ascribed to another local energy minimum of the system. As elucidated by Whyman et al., [39] considering a droplet with a constant volume, the total interfacial energy E of the system is a function of the contact angle θ, and its derivative is expressed as: ...
Article
Full-text available
When a multi‐component fluid contacts arigid solid substrate, the van der Waals interaction between fluids and substrate induces a depletion/adsorption layer depending on the intrinsic wettability of the system. In this study, we investigate the depletion/adsorption behaviors of A−B fluid system. We derive analytical expressions for the equilibrium layer thickness and the equilibrium composition distribution near the solid wall, based on the theories of de Gennes and Cahn. Our derivation is verified through phase‐field simulations, wherein the substrate wettability, A−B interfacial tension, and temperature are systematically varied. Our findings underscore two pivotal mechanisms governing the equilibrium layer thickness. With an increase in the wall free energy, the substrate wettability dominates the layer formation, aligning with de Gennes’ theory. When the interfacial tension increases, or temperature rises, the layer formation is determined by the A−B interactions, obeying Cahn's theory. Additionally, we extend our study to non‐equilibrium systems where the initial composition deviates from the binodal line. Notably, macroscopic depletion/adsorption layers form on the substrate, which are significantly thicker than the equilibrium microscopic layers. This macroscopic layer formation is attributed to the interplay of phase separation and Ostwald ripening. We anticipate that the present finding could deepen our knowledge on the depletion/adsorption behaviors of immiscible fluids.
... Meanwhile, the deposited nanospheres significantly improve the surface roughness of the modified membrane, that is approximately 2.55 times higher than that of the original membrane (see figures 3(h)-(i)), so the contact area between oil droplets and the membrane surface is significantly reduced, associated with more water molecules being embedded in the concave and convex pores. Due to the enhanced surface roughness, the wetting status of oil droplets on the membrane surface further changes from Wenzel state to Cassie state, leading to a high underwater oil contact angle of 180° [48][49][50]. ...
Article
Full-text available
Membrane distillation technology could utilize low-grade heat to desalinate brine, but the membrane material often suffers from disadvantages of low permeation flux and weak robustness to contaminants. To address these issues, the commercial polytetrafluoroethylene (PTFE) membrane was modified by cost-effective chemicals of tannic acid and (3-Aminopropyl)-triethoxysilane (APTES) to construct hydrophilic/underwater superoleophobic nano-rough structures on the surface to enhance its flux and oil-fouling resistance in direct contact membrane distillation. The results show that a high underwater oil contact angle of 180° is observed to the membrane surface due to the rough nanostructures functionalized by abundant hydroxyl groups. Despite the additional mass transfer resistance provided by the rough nanostructures, the flux was increased noticeably. This is mainly attributed to the strong interactions between the abundant hydroxyl groups of hydrophilic layer surface and water molecules, leading to a part of free water staying at intermediate transition state (IW). The mass transfer resistance of the hydrophilic layer itself is reduced as a consequence of decreased evaporation enthalpy of water, thereby increasing the flux. Moreover, while the flux of the pristine membrane is reduced by 84.18%, the flux of Janus membrane remains the same when treating mineral oil brine emulsions with oil concentration up to 1500 ppm in comparison with the result for 35 g l⁻¹ brine solution, indicating that the Janus membrane is safe from the oil contamination. Our work provides a fine guidance for membrane distillation to treat high oily brine.
... Combining these two models is recommended to describe contact angle behavior across a wide velocity range [351]. Additionally, several contact angle hysteresis models have been developed to account for contact angle changes under complex conditions [347,352]. ...
Article
Boiling is an efficient heat transfer mode with significant potential for thermal management in high-power electronic equipment. However, a comprehensive understanding of the boiling process, which encompasses bubble nucleation, growth, coalescence, slipping, and detachment across various scales, remains challenging. Molecular dynamics simulation, lattice Boltzmann, and computational fluid dynamics methods are popular and powerful tools for investigating boiling heat transfer phenomena at microscopic, mesoscopic, and macroscopic scales. These methods enable researchers to uncover the underlying boiling mechanisms and propose heat transfer enhancement techniques. Therefore, this paper provides a comprehensive review of boiling heat transfer, spanning from atomistic to mesoscopic and macroscopic scales, utilizing these three numerical methods. It addresses critical issues related to nanoscale bubble nucleation mechanisms, pool boiling, and flow boiling, and proposes potential solutions and future researches, supplementing our previous review [Some advances in numerical simulations of multiscale heat transfer problems and particularly for boiling heat transfer, Annu. Rev. Heat Transf., 6 (2022) 217-269]. Besides, by shedding light on the characteristics of these numerical methods in studying boiling heat transfer, this paper aims to foster their development and advance enhanced heat transfer technologies.
... This discrepancy presumably arises because the real-world droplet detachment differs considerably from the idealized scenario. In the latter case, droplets with a constant volume (V 0 ), an intrinsic contact angle (θ e ), and spherical sections with curved (liquid-vapor) and flat (liquid-solid) areas (A C and A P , respectively) obey the geometric relationship 28,40,45,46 . This is only possible when the apparent contact angle does not change significantly from its value at thermodynamic equilibrium. ...
Article
Full-text available
Liquid droplets hanging from solid surfaces are commonplace, but their physics is complex. Examples include dew or raindrops hanging onto wires or droplets accumulating onto a cover placed over warm food or windshields. In these scenarios, determining the force of detachment is crucial to rationally design technologies. Despite much research, a quantitative theoretical framework for detachment force remains elusive. In response, we interrogated the elemental droplet–surface system via comprehensive laboratory and computational experiments. The results reveal that the Young–Laplace equation can be utilized to accurately predict the droplet detachment force. When challenged against experiments with liquids of varying properties and droplet sizes, detaching from smooth and microtextured surfaces of wetting and non-wetting chemical make-ups, the predictions were in an excellent quantitative agreement. This study advances the current understanding of droplet physics and will contribute to the rational development of technologies.
Article
This article presents an experimental study about boundary slippage on the film thickness of hydrodynamic lubrication (HL) using a custom‐made slider‐on‐disc bearing testing apparatus. The interfaces with different affinity were obtained by surface energy modification of sliders with various oleophobic coatings, which are characterised by their contact angle (CA) and contact angle hysteresis (CAH). To study the mechanism of interfacial slip on HL under different wettability constraints, the film thickness and velocity profiles under shear were measured using interference and fluorescence photobleached method, respectively. The results showed that the CAH could better characterise the influence of interface effect on the film thickness of HL, which was explained by the correlation between CAH and the interface potential barrier. Furthermore, it was found that the slip velocity increased with lubricant viscosity and shear rate, which can be explained by the spatial heterogeneity of the flow in conformal contact and the critical shear stress slip model.
Article
Titania nanotubes fabricated over grade V medical Ti6Al4V alloys (Ti64) in fresh pure electrolytes EG and EGD in order to meet the requirements of a significant implant material for biomedical application. Wettability and contact angle hysteresis is a matter of concern for the nano-textured surface that needs to be understand for its biological activity in correlation to surface properties. Therefore, titania nanotubes have been fabricated in two different electrolytes to understand the wetting behaviour. The present study shows an increment in the wettability from hydrophilic to hydrophobic compared to emery-polished Ti-Alloy. Topographical change of TiO2 nanotubes has also been observed in two different electrolytes. Additionally, change in the wall thickness has been identified in the TiO2 nano fabricated on fresh electrolytes.
Article
A strict and generalized definition of the concept of disjoining pressure of thin layers (of liquid, gas, vacuum) is presented as a difference between the normal component of the tensor pressure in a layer and the pressure in a phase from which the layer has been formed by thinning out. The statics and the kinetics of systems including liquid interlayers of nonuniform thickness are considered.
Article
Contact angle relaxation was measured for captive air bubbles placed on solid surfaces of varying degrees of heterogeneity, roughness, and stability, in water. The experimental results indicate that both advancing and receding contact angles undergo slow relaxation in these water-air-solid systems, due to instabilities of the three-phase contact line region. It is shown that the advancing contact angle decreases and the receding contact angle increases for many systems over a period of a few hours. Also, examples of reverse progressions are reported. Additionally, in extreme cases, the contact angle oscillates down and up, over and over again, preventing the system from stabilization/equilibration. Four different mechanisms are proposed to explain the contact angle relaxation. These include (i) pinning of the three-phase contact line and its slow evolution; (ii) the formation of microdroplets on the solid surface and their coalescence with the base of the gas bubble, which causes dynamic behavior of the three-phase contact line; (iii) deformation of the solid surface and its effect on the apparent contact angle; and (iv) chemical instability of the solid.
Article
In this paper, a method to measure the advancing and receding contact angles on individual colloidal spheres is described. For this purpose, the microspheres were attached to atomic force microscope cantilevers. Then the distance to which the microsphere jumps into its equilibrium position at the air-liquid interface of a drop or an air bubble was measured. From these distances the contact angles were calculated. To test the method, experiments were done with silanized silica spheres (4.1 μm in diameter). From the experiments with drops, an advancing contact angle of 101 ± 4° was determined. A receding contact angle of 101 ± 2° was calculated from the jump-in distance into a bubble. Both experimental techniques gave the same contact angle. In contrast, on similarly prepared planar silica surfaces, a clear hysteresis was measured with the sessile drop method; contact angles of 104.5 ± 1° and 93.8 ± 1° were determined for the advancing and receding contact angles, respectively.
Article
Wetting of SiO2/Si/SiO2 slides by phospholipid dispersions has been determined over a range of phospholipid concentrations and times of interaction between vesicles and the solid surface. Both advancing and receding dynamic contact angles () increased and the contact angle hysteresis decreased as a function of the phospholipid concentration (or interaction time), attaining a maximum (a minimum). Maximization of both the advancing and receding contact angles and minimization of the contact angle hysteresis are associated with an increase in the chemical homogeneity of the surface and one-bilayer deposition. Determination of contact angles is a Powerful, quick, and simple technique to establish experimental conditions for bilayer deposition on solid surfaces in general. This result may be of importance for further advancements in rapidly developing research areas such as the design of biosensors, immunodiagnosis, and the development of biocompatible materials.
Article
It has already been asserted, by Mr. Monge and others, that the phenomena of capillary tubes are referable to the cohesive attraction of the superficial particles only of the fluids em­ployed, and that the surfaces must consequently be formed into curves of the nature of lintearias, which are supposed to be the results of a uniform tension of a surface, resisting the pressure of a fluid, either uniform, or varying according to a given law. Segner, who appears to have been the first that maintained a similar opinion, has shown in what manner the principle may be deduced from the doctrine of attraction, but his demonstration is complicated, and not perfectly satisfactory; and in applying the law to the forms of drops, he has neglected to consider the very material effects of the double curvature, which is evidently the cause of the want of a perfect coinci­dence of some of his experiments with his theory. Since the time of Segner, little has been done in investigating accurately and in detail the various consequences of the principle. It will perhaps be most agreeable to the experimental phi­losopher, although less consistent with the strict course of logical argument, to proceed in the first place to the comparison of this theory with the phenomena, and to inquire afterwards for its foundation in the ultimate properties of matter. But it is necessary to premise one observation, which appears to be new, and which is equally consistent with theory and with experiment; that is, that for each combination of a solid and a fluid, there is an appropriate angle of contact between the surfaces of the fluid, exposed to the air, and to the solid. This angle, for glass and water, and in all cases where a solid is perfectly wetted by a fluid, is evanescent: for glass and mer­cury, it is about 140°, in common temperatures, and when the mercury is moderately clean.
Article
The relation between drop radius, r, the force to move the three phase contact line and the advancing and receding contact angles θA and θR is studied. To keep the line energy (energy per 2πr, also named line tension) independent of r, the modified Young equation predicts that the advancing and receding contact angles, θA and θR, change considerably with r. As shown by many investigators, θA and θR change negligibly, if at all, with r. We quantify recent evidences showing that the line energy is a function of the Laplace pressure and show that this way the modified Young equation is correct and still θA and θR should hardly change with r. According to our model, the small surface deformation associated with the unsatisfied normal component of the Young equation results in higher intermolecular interactions at the three phase contact line which corresponds to a higher retention force. This time increasing effect is supported by recent experiments.