• No results found

Capillary bridges between spherical particles under suction control: Rupture distances and capillary forces

N/A
N/A
Protected

Academic year: 2021

Share "Capillary bridges between spherical particles under suction control: Rupture distances and capillary forces"

Copied!
13
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Capillary bridges between spherical particles under suction control:

Rupture distances and capillary forces

Chao-Fa Zhao

a,b

, Niels P. Kruyt

a,∗

, Olivier Millet

b a

Department of Mechanical Engineering, University of Twente, P.O. Box 217, Enschede, 7500 AE, the Netherlands bLaSIE-UMR CNRS, Université de La Rochelle, 23 Avenue Albert Einstein, La Rochelle, 17000, France

a b s t r a c t

a r t i c l e i n f o

Article history: Received 5 August 2019

Received in revised form 23 September 2019 Accepted 29 September 2019

Available online 18 October 2019

This study focuses on properties of axisymmetric capillary bridges between spherical particles in the pendular re-gime under suction control. Using the toroidal approximation in combination with the governing Young-Laplace equation, analytical expressions for the rupture distance (dependent on the suction) and for the capillary force (dependent on the suction and the interparticle separation distance) have been obtained that do not involve any calibrated coefficient. The developed analytical expressions are effective for values of the dimensionless suction larger than 104. To predict capillary forces and rupture distances for a wider range of values of the

dimen-sionless suction, closed-form expressions for rupture distances and capillary forces have been obtained by curve-fitting to a large dataset of numerical solutions of the Young-Laplace equation. It is shown that these expressions can also be employed for capillary bridges between spheres with unequal radii when the Derjaguin radius is employed.

© 2019 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http:// creativecommons.org/licenses/by-nc-nd/4.0/). Keywords: Capillary bridge Suction control Rupture distance Capillary force Young-Laplace equation Toroidal approximation 1. Introduction

Different types of cohesive forces, such as van der Waals forces [1], electrostatic forces [2] and capillary forces [3] occur in granular mate-rials. For wet granular materials, capillary forces are larger than the other cohesive forces.

Hence, cohesion in wet granular materials originates from capillary forces between the constituting particles [4–8]. With decreasing liquid content, various regimes are encountered [7,9–12]: (1) the capillary re-gime where the particles are fully immersed in liquid, (2) the funicular regime with coalesced capillary bridges, with liquid clusters that are si-multaneously in contact with at least three particles and (3) the pendu-lar regime where a capilpendu-lary bridge is associated with exactly two particles.

The current study focuses on capillary bridges in the pendular re-gime and situations where the surface of the capillary bridge is axisym-metric. The effect of viscosity is not considered here, since the capillary bridges are assumed to be in quasi-static equilibrium. The influence of gravity on capillary bridges that is investigated in Ref. [13] is neglected in the present study.

Important properties of capillary bridges between two spheres in the pendular regime are the rupture distance (i.e. the maximum

separation distance between the particles for which a capillary bridge exists) and the capillary force. The dependence of the rupture distance on the volume of the capillary bridge and of the capillary force on the volume of the capillary bridge and on the separation distance has been extensively investigated [3,14–18]. This type of problem is called volume control here.

Besides properties of capillary bridges under volume control (where the volume is kept constant), such properties under suction control (where the pressure difference between the surrounding gas and the liquid inside the capillary bridge is kept constant) are also important. For instance, in Discrete Element Method [19] simulations of the behav-iour of wet granular materials, expressions for given capillary bridge volume are appropriate for undrained tests, whereas expressions for given suction pressure are appropriate for drained tests [20–23]. In practice, the behaviour of wet granular materials under suction control is of great interest when the humidity of the surrounding gas is con-trolled. The relationship between suction and humidity under condi-tions of thermodynamic equilibrium is discussed in detail in Section2.5. The current study focuses on properties of capillary bridges under suction control.

Properties of capillary bridges are determined by the capillary bridge surface geometry, which is described by the well-known Young-Laplace equation [3,14–17,24–26], according to which the mean curvature of the surface is constant. It has been shown experimentally that the Young-Laplace equation accurately describes the properties of capillary bridges [27–35]. According to the Young-Laplace equation [14–18,24], ∗ Corresponding author.

E-mail addresses:c.zhao-1@utwente.nl(C.-F. Zhao),n.p.kruyt@utwente.nl(N.P. Kruyt), olivier.millet@univ-lr.fr(O. Millet).

https://doi.org/10.1016/j.powtec.2019.09.093

0032-5910/© 2019 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/). Contents lists available atScienceDirect

Powder Technology

(2)

the mean curvature of the capillary bridge surface H is proportional to the value of suctionΔp.

H¼ Δp γ ¼

pext−pint

γ H

¼ H  R ; ð1Þ

where pextand pintare the pressures outside and inside the capillary

bridge;γ is the surface tension between water and gas ( γ ¼ 72:8 10−3N/m for water-air interfaces at a temperature of 20C); H is

made dimensionless using the Derjaguin radius R‾ [18,36], giving the di-mensionless suction H. The (effective) Derjaguin radius R corresponding to a small spherical particle with radius Rsand a large spherical particle

with radius RLis defined by

R¼ 2RsRL

Rsþ RL :

ð2Þ Properties of capillary bridges under suction control have been stud-ied using the following complimentary approaches: numerically [20,22,30,37–39], analytically [37,38] and experimentally [30]. Numeri-cal solutions of the nonlinear Young-Laplace equation are effectively exact, but are cumbersome and time-consuming when used a large number of times. Analytical methods have a clear physical meaning, but they have a limited range of validity as they involve some approxi-mations. Experiments directly provide information on the capillary bridge properties, but they are time-consuming and difficult to perform due to the effects of gravity and particle surface roughness. To the authors’ best knowledge, explicit expressions for the capillary force under suction control have not been reported in the literature.

According to analyses of numerical solutions to the Young-Laplace equation, the relation between rupture distance and capillary bridge volume under suction control is very different from that under volume control [22,37–40]. Numerical solutions also show that the capillary force decreases with increasing value of suction (for the same separa-tion distance).

Analytical theories are generally based on the toroidal approxima-tion [3,16,37,38,40,41]. In this approximation the meridional profile of capillary bridges is represented by part of a circle. Under volume con-trol, expressions for the capillary force have been reported [37,38,40,41]. Recently, analytical theories based on the ellipse approx-imation [14,15,18] have been developed for volume control. In this ap-proximation, the meridional profile of capillary bridges is represented by part of an ellipse, and the Young-Laplace equation has been taken into account, contrary to the toroidal approximation. For given capillary bridge volume and separation distance, expressions for the rupture dis-tance and the capillary force have been derived, in which no calibrated coefficients are required.

Gras et al. [30] have experimentally measured capillary forces and rupture distances of capillary bridges between two equal-sized spheri-cal particles under suction control. Their experimentalfindings are con-sistent with those obtained from numerical solutions to the Young-Laplace equation [20,22,30,37–39]. They have also suggested a closed-form expression for the rupture distance, based on their experimental results and numerical solutions to the Young-Laplace equation. How-ever, the dimensionless suction in their closed-form expression for the rupture distance is limited to small values (0≤H≤15).

For capillary bridges under volume control or suction control, it has been numerically shown that two solutions to the Young-Laplace equa-tion may exist [14–16,18,22,37]. To select the physically-relevant branch (also referred to as the stable solution), various criteria have been developed in Refs. [22,25,42]. However, the effectiveness of these criteria has not been verified for both volume and suction control conditions.

In experimental studies with wet granular materials, it is not possi-ble to continually increase the liquid pressure to large values while keeping the gas pressure at atmospheric level, since the liquid may

vaporize under high (negative) pressure levels [43,44]. Therefore, the axis translation technique that is suitable for lower values of suction and the vapour equilibrium method that is applicable for higher values of suction have been developed. In the vapour equilibrium method the relative humidity of the gas is controlled. As described in Refs. [44,45], the vapour equilibrium technique cannot accurately control the dimen-sionless suction when the relative humidity h is close to 100%, due to the strongly nonlinear h−Hcurve that requires a high-precision humidity

sensor. The relation between the relative humidity h and the dimen-sionless suction H is generally given by the Kelvin equation [37,38,46]. Therefore, it is valuable to (also) formulate the developed expressions for the rupture distance and the capillary force in terms of humidity and separation distance.

The range of dimensionless suctionHof interest in the current study is H≥1. ForHb1, the capillary force is relatively small and hence it is not

important. For granular soils, the mean particle size is about 7:4  10−5 m [44]; accordingly, the dimensionless suction His approximately larger than 102(H N 102). The values of Hconsidered here have a much larger range than that in Ref. [30], in which values 0≤H≤15

have been considered.

In the present study, properties of capillary bridges under suction control are determined using the complimentary approaches of an ana-lytical theory as well as analyses of numerical solutions to the governing Young-Laplace equation. The analytical theory is valid for large values of the dimensionless suction (H≥104

) that correspond to small capillary bridge volumes. A large dataset of numerical solutions to the Young-Laplace equation is used to formulate closed-form expressions (by curve-fitting) that are valid for a wider range of dimensionless suction H≥1, and thus both small and large capillary bridge volumes have been taken into account. Specific objectives of this study are to: • demonstrate the effectiveness of the surface free-energy based

crite-rion for selecting the physically-relevant solutions to the governing Young-Laplace equation (obtained by a high-resolution integration method) under both suction and volume control conditions; • develop an analytical theory for the rupture distance and the capillary

force under suction control, by employing the toroidal approximation; • formulate closed-form expressions for the rupture distance and the capillary force under suction control by curve-fitting to a large dataset of numerical solutions of the governing Young-Laplace equation; • consider the influence of the particle size ratio on the rupture distance

and the capillary force under suction control by using the Derjaguin radius (which is valid for small capillary bridge volumes).

The outline of this study is as follows. Section2gives the description of the geometry of the capillary bridge, the numerical solution method for the governing Young-Laplace equation, the criterion for selecting the physically-relevant solutions and the vapour equilibrium method for suction control. In Section3analytical expressions are derived for the rupture distance and the capillary force for given suction and sepa-ration distance, based on the toroidal approximation. In Section4

closed-form expressions are given for the capillary force and the rupture distance by curve-fitting to the large dataset of numerical solutions to the Young-Laplace equation. Section5develops the obtained expres-sions for unequal-sized spherical particles by using the Derjaguin ra-dius. Section 6 summarizes the contributions of this study and discusses further developments.

2. Capillary bridge geometry and numerical solution of the Young-Laplace equation

In this Section, the geometry of the capillary bridge is described. The Young-Laplace equation is given that describes the geometry of the cap-illary bridge surface. The numerical solution method for the Young-Laplace equation is presented. The surface free-energy criterion for

(3)

selecting physically-relevant branches of the numerical solutions is then discussed, for both suction and volume control conditions. The method for controlling suction by the vapour equilibrium method using the Kel-vin equation is also described.

2.1. Geometry of the capillary bridge

An axially symmetric capillary bridge between two spherical parti-cles with radii Rsand RL(RL≥Rs) is considered. The subscripts s and L

refer to properties of the small and large particle, respectively (conform the notation in Ref. [18]).

As shown inFig. 1, the axial coordinate is denoted by X, and the me-ridional profile of the capillary bridge is described by Y Xð Þ. The neck where the radius of the capillary bridge is smallest is taken at X¼ 0. The neck radius is denoted by Y0. Hence

Y 0ð Þ ¼ Y0 Y0ð Þ ¼ 0 ;0 ð3Þ

where the slope of the meridional profile Y Xð Þ is denoted by Y0ð Þ ¼ dYX

=dX.

The meridional coordinates Xð cs; YcsÞ and Xð cL; YcLÞ of the three-phase

contact circles at the small and large spheres are related to separation distances (Ssand SL), and half-filling angles (δsandδL) (seeFig. 1) by

Ref. [18].

Xcs¼ Ssþ Rsð1−cos δsÞ Ycs¼ Rssinδs

XcL¼ −SL−RLð1−cos δLÞ YcL¼ RLsinδL: ð4Þ

For small capillary bridge volumes, the contact radii Ycsand YcLare

very small. Therefore, Eq.(4)can be approximated by

Xcs≅Ssþ 1 2 Y2 cs Rs Ycs≅Rsδs XcL≅−SL− 1 2 Y2 cL RL YcL≅RLδL: ð5Þ

In addition, the slopes of the meridional profiles at the contact circles in-volve the contact angleθ (see alsoFig. 1). These slopes are determined by Ref. [18]. Y0ðXcsÞ ¼ cot δð sþ θÞ ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi R2 s−Y2cs q −TYcs Ycsþ T ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi R2 s−Y 2 cs q ≡ Y0cs Y0ðXcLÞ ¼ −cot δð Lþ θÞ ¼ − ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi R2 L−Y 2 cL q −TYcL YcLþ T ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi R2L−Y 2 cL q ≡ Y0cL; ð6Þ

where T is a short-hand notation for the term

T¼ tan θ : ð7Þ

Here the contact angleθ is in radians. 2.2. Young-Laplace equation and capillary force

It is well-known that the geometry of the capillary bridge surface is described by the Young-Laplace equation [14–18,24], according to which the mean curvature of the surface is constant. For the axially-symmetric capillary bridges considered in this study, the Young-Laplace equation can be expressed as

K½Y Xð Þ ≡ Y″ 1þ Y02  3=2− 1 Y 1 1þ Y02  1=2¼ H ; ð8Þ

where H is given by Eq.(1)and K½Y is the mean curvature operator; Y0¼ dY=dX and Y″¼ d2

Y=dX2

. For this nonlinear second-order ordi-nary differential equation, boundary conditions are given by Eqs.(3), (4) and (6).

The Young-Laplace equation 8ð Þ has a first integral Λ½Y Xð Þ ≡ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiY 1þ Y02 q þ1 2HY 2 ¼ λ ; ð9Þ

whereλ is a constant. This expression corresponds to axial force equilib-rium along the capillary bridge [14,16,25]. By substitution of Eqs.(3) and (6)into Eq.(9), it follows that the mean curvature H and the constantλ are given by

Y0− YcL ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1þ Y0 cL 2 q Y2cL−Y 2 0 ¼1 2H¼ Y0− Ycs ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1þ Y0 cs 2 q Y2cs−Y 2 0 λ ¼ Y0þ 1 2HY 2 0: ð10Þ

The capillary force Fcapbetween two spherical particles is then

deter-mined by

Fcap¼ λ 2πγð Þ ; ð11Þ

where Fcapandλ are both dimensional (contrary to in Refs. [14,15]).

2.3. Numerical solutions of the Young-Laplace equation

When values of the contact radius of the small particle Ycsand the

neck radius Y0are specified, the meridional profile X Yð Þ, the dual of Y

X

ð Þ , is obtained by using a high-resolution numerical integration method, as described in Ref. [18]. The capillary bridge volume (for each particle, i.e. the part up to the neck located at X¼ 0) is evaluated by numerical integration from the meridional profileY Xð Þ. The total cap-illary bridge volume V and total separation distance 2S are determined by the capillary bridge volumes and the separation distances of each particle

V¼ Vsþ VL 2S¼ Ssþ SL; ð12Þ

where Vsand VLare the capillary bridge volumes for the small and large

particle respectively; Ssand SLare the separation distances for the small

and large particle respectively, as shown inFig. 1.

Zhao et al. [18] compiled a database consisting of a large number of numerical solutions to the Young-Laplace equation, for various values of half-filling angles δs≤60∘, ratios of neck radius to contact radius 0≤Y0=

Ycs≤1 and size ratios of RL=Rs = 1, 2, 8 and 128 and contact angles of

θ = 0∘, 20and 40. For each half-filling angle δ

s and ratio Y0=Ycs,

the resulting separation distances (Ss, SLand S), capillary bridge volumes

Fig. 1. Meridional geometry of a capillary bridge between two unequal-sized spheres RL≥Rs

ð Þ with contact angle θ. The meridional profile is described by Y Xð Þ; the neck radius is denoted as Y0; the meridional coordinates of the contact circles of the small and the large particles are Xð cs; YcsÞ and Xð cL; YcLÞ, respectively; the half-filling angles are δsandδL, respectively; the separation distances, with respect to the neck located at X¼ 0, for the small and the large particle are Ssand SL, respectively. The total separation distance between the two particles is 2S. (Figure from Ref. [18]).

(4)

(Vs, VLand V), capillary force (λ) and mean curvature (H) have been

stored. For any combination of mean curvature and separation distance H; S

ð Þ of interest, the (scaled) capillary force λ is obtained by interpolat-ing the dataset, usinterpolat-ing the method described in Ref. [14]. The same data-base and interpolation method have been employed here.

2.4. Stable branches of solutions to the Young-Laplace equation

For specified values of the neck radius Y0and contact radius Yc, the

solution for the meridional profile Y Xð Þ, as determined by direction in-tegration (see for instance Eq.(13)in Ref. [18]), is unique. However, this may not be the case with specified volume V and separation S (so under volume control) [14–16,18] or with specified suction H and sep-aration S (so under suction control) [22].

In general, only a single branch of such solutions is stable and is ob-served experimentally. Criteria based on the minimum of some energy have been proposed to identify the stable branch of solutions [14–16,22,25,47,48].

The energy considered in Refs. [16,47,48] is the total surface energy of the system Esurf, given by the sum of surface areas weighted by the

corresponding interfacial energies per area (i.e. surface tensions)

Esurf¼ γAlgþ γlsAlsþ γsgAsg: ð13Þ

Hereγ,γlsandγsgare the surface tensions for liquid-gas, liquid-solid and

solid-gas interfaces, and the corresponding interfacial surface areas are Alg, Alsand Asgrespectively.

An energy that also involves other sources of free energy has been considered in Ref. [22].

EsurfþΔpV¼ Esurfþ ΔpV : ð14Þ

whereΔp is the pressure difference between gas and liquid; V is the cap-illary bridge volume.

Besides the energies in Eqs.(13) and (14), the energy considered by Ref. [25] is equal to that in Eq.(13)but shifted by the constant 4πγR‾2 and the energy EsurfþpextV¼ Esurfþ paV in Ref. [42] does not involve

the pressure differenceΔp that determines the mean curvature in the Young-Laplace equation(8).

Under volume control (withfixed capillary bridge volume V), Lian et al. [16] have demonstrated that the criterion of minimal Esurfis

effec-tive in identifying the stable branch of solutions of the Young-Laplace equation. They also found that the minimum free surface energy crite-rion is equivalent to the minimum surface energy of the liquid-gas in-terface, i.e.γAlg. However, for capillary bridges under suction control,

the validity of the criterion of a minimum of the energy defined in Eq.(13)has not been investigated in literature.

Under suction control (withfixed suction H), Duriez and Wan [22] have shown that the criterion of minimal EsurfþΔpVis effective in

identi-fying the stable branch of solutions of the Young-Laplace equation. However, the appropriateness of this criterion for capillary bridges under volume control has not been investigated.

The suitability of criteria based on Esurf and EsurfþΔpVfor capillary

bridges under volume as well as suction control, respectively, has been investigated inAppendix A. There it is shown that the criterion of a minimum of the surface free energy defined by Eq.(13)is appropri-ate for identifying the stable branch of solutions to the Young-Laplace equation, under both volume and suction control conditions.

2.5. Relationship between dimensionless suction and relative humidity The suction in wet granular materials can be controlled through sev-eral techniques. Among these methods, the vapour equilibrium method has been extensively employed for unsaturated soils [44,45]. With this method, the suction is controlled by varying the vapour pressure,

which is equivalent to changing of the relative humidity. The theoretical basis of this method is provided by the Kelvin equation [37,38,46,49] that relates vapour pressure pvand saturation vapour pressure psatto the

di-mensionless suction Hunder conditions of thermodynamic equilibrium ln pv psat  ¼ −Δp ρlRvTt ¼ −γH ρlRvTtR;  ð15Þ Here the ratio pv=psatis defined as the relative humidity h;ρlis the density

of liquid in kg=m3; R

vis the (specific) gas constant of liquid vapour (Rv¼

461:5 J=kgK for water vapour) and Ttis the absolute temperature in K.

Accordingly, the dimensionless suction Hcan be related to the relative humidity h by

H¼ −ρlRvTt

γ R‾ln h : ð16Þ

The relationship between dimensionless suction Hand relative humid-ity h is illustrated inFig. 10of [37].

The large dataset of numerical solutions to the Young-Laplace equa-tion described in Secequa-tion2.3can also be used to study properties of cap-illary bridges under humidity control, by employing Eq. (16). In addition, the effect of temperature Tton the dimensionless suction H

at a prescribed humidity level can also be investigated. By using Eq.(16), the dimensionless suction Hpresent in closed-form expres-sions in the following Sections can be converted to terms involving the relative humidity h.

3. Analytical expressions for rupture distances and capillary forces In this Section, capillary bridges between equal-sized spherical parti-cles under suction control are considered. The meridional profile of the capillary bridge is represented by part of a circle, i.e. the toroidal approx-imation is employed [3,37,38,40,41]. Based on the toroidal approxima-tion, analytical expressions for the rupture distance and the capillary force are derived in Sections3.2 and 3.3, respectively. The accuracy of these expressions is demonstrated by comparisons with numerical so-lutions of the Young-Laplace equation.

3.1. Toroidal approximation

In the toroidal approximation, part of a circle is used to represent the meridional profile of the capillary bridge, see for instance Refs. [3,16,37,38,40,41]. In earlier developments [3,16,37,38,40,41], cap-illary forces were expressed as functions of separation distance and half-filling angle. It is therefore valuable to further develop this approxima-tion in order to directly calculate rupture distances and capillary forces for given values of suction and separation distance.

Numerical solutions to the Young-Laplace equation can be represented in dimensionless form using the scaling method suggested in Refs. [14,15,18]. For equal-sized spherical particles, the capillary bridge is symmetric with respect to the neck located at X¼ 0, thus Xcs¼

− XcLand Ycs¼ YcL. Therefore, the scaling method is expressed as

P¼jXcij

Yci

Q¼Y0

Yci:

ð17Þ As discussed in Refs. [14,15,18], the dimensionless coordinate P repre-sents the scaled axial position of the contact circle and the dimension-less coordinate Q represents the scaled neck radius.

The‘closure’ relationship P ¼ F Qð Þ between the scaled coordinates P and Q in Eq.(17)according to the toroidal approximation is given by Refs. [15,18]. P¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 q þ T   1−Q ð Þ : ð18Þ

(5)

Under volume control (so forfixed value of the capillary bridge vol-ume), it has been found that the toroidal approximation Eq.(18)is ac-curate for small separation distances, but not for large separation distances [14,15,18,28]. For suction control,Fig. 2shows the comparison between the relationship between the scaled coordinates P and Q from numerical solutions of the Young-Laplace equation and according to the toroidal approximation, for contact anglesθ of 0∘and 40∘. This Fig-ure shows that, under suction control, the toroidal approximation well represents numerical solutions to the Young-Laplace equation, both for small and large separation distances, for the stable branch of solutions.

Under volume control, the critical point Qð ; PÞ (separating the

sta-ble and the unstasta-ble branch) of the relation P¼ F Qð Þ corresponds to the maximum value of P [14,15,18]. However, the P−Q relation for suction control does not have such a property. Therefore, it is not necessary to employ the ellipse approximation [14,15,18] (that has a peak value of P) for capillary bridges under suction control.

For capillary bridges between equal-sized spherical particles, Rs¼ RL

¼ R, and hence S¼S R¼ Ss R¼ SL R Y  c¼ Ycs R ¼ YcL R Y  0¼ Y0 R H¼ H  R λ¼ λ R V ¼V R3: ð19Þ

These dimensionless variables have been adopted in the following der-ivations for the rupture distance and the capillary force.

3.2. Rupture distance

Here it is shown that the critical point Qð ; PÞ (indicated inFig. 2)

that separates the stable and unstable branch of solutions can be well predicted by the toroidal approximation, and hence the rupture dis-tance under suction control can be estimated by the toroidal approximation.

For capillary bridges between equal-sized particles with small vol-umes, it follows from Eqs.(5) and (17)that P can be expressed in terms of separation distance Sand contact radius Yc.

P¼S Yc þ1 2Y  c: ð20Þ

Substituting Eqs.(17) and (20)into Eq.(18), it follows that the separa-tion distance Scan be expressed in terms of the contact radius Ycand

the neck radius Y0.

S¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 q þ T   Yc− 1 2Y 2 c− ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 q þ T   Y0: ð21Þ

Using Eq.(10), the dimensionless neck radius Y0can be expressed in

terms of the dimensionless contact radius Ycand the dimensionless

suc-tion H Y0¼ − 1 Hþ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2Yc Tþ Yc   HpffiffiffiffiffiffiffiffiffiffiffiffiffiffiT2þ 1þ 1 H2þ Y 2 c : s ð22Þ Substituting Eq.(22)into Eq.(21), the separation distance Sis obtained as a function of contact radius Ycand dimensionless suction H:

S¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 q þ T   Yc1 2Y 2 c− ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 q þ T    −H1þ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2Yc Tþ Yc   HpffiffiffiffiffiffiffiffiffiffiffiffiffiffiT2þ 1þ 1 H2þ Y 2 c s ! : ð23Þ

Eq.(23)can be rewritten as a quartic equation in Yc.

K4Y4c þ K3Y3c þ K2Y2c þ K1Ycþ K0¼ 0 ; ð24Þ

where the coefficients K4, K3, K2, K1and K0(functions of the

dimension-less suction H, the dimensionless separation distance Sand the contact angleθ) are given in Appendix B. For given suction and separation dis-tance, Eq.(24)has two real and positive solutions for Ycthat are

phys-ically relevant. At the critical separation distance 2Scrit, these two

solutions are the same. This occurs when the discriminant of the quartic equation 24ð Þ equals zero [50]. Based on this property, the rupture dis-tance 2Scritcan be expressed as

2Scrit¼ 2 H 1 ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 p  1− 27 32þ 27 16Tþ 81 64T 21 H 1=3) : ( ð25Þ The detailed derivation of Eq.(25)is given in Appendix B. With relation-ship Eq.(25), the critical separation distance 2Scritof capillary bridges

can be determined for specified suction values H(and contact angleθ).

Fig. 2. Relation between scaled coordinates P and Q (defined in Eq.(17)) from numerical solutions to the Young-Laplace equation and predictions according to the toroidal approximation for dimensionless suction H≥104. The critical point Qð ; PÞ is indicated that separates the stable and unstable branches. (a) Contact angle θ ¼ 0; (b) contact angleθ ¼ 40.

(6)

Fig. 3shows comparisons of the critical separation distance 2Scrit ex-tracted from the large database of numerical solutions of the Young-Laplace equation and that predicted by the rupture criterion of Eq.(25). Note that the product of rupture distance and dimensionless suction is compared, as that more clearly shows its accuracy (as the rup-ture distance itself is proportional to 1=H, which varies by multiple

or-ders of magnitude). The analytical rupture criterion agrees very well with the numerical solutions of the Young-Laplace equation, for contact angles 0∘≤θ≤40∘. When H≅104, the deviations are smaller than 2% for

θ ¼ 0∘, 1% forθ ¼ 20∘and 0:5% for θ ¼ 40∘. For larger values of H, the de-viations are much smaller, not more than 0:5% when H≥106

. Note that this analytical result does not involve any calibratedfitting coefficient. 3.3. Capillary force

The dimensionless capillary forceλcan be determined, for given

values of mean curvature Hand neck radius Y0, through the second

ex-pression of Eq.(10). For capillary bridges under suction control, the neck radius Y0is a function of the dimensionless suction Hand the

dimen-sionless separation distance S, i.e. Y0¼ Y0ðH; SÞ . Therefore, for

given values of dimensionless suction Hand separation distance S, the capillary force between two equal-sized spherical particles can be determined by λ¼ Y 0 H; S  ð Þ þ1 2H Y2 0 H; S  ð Þ ; ð26Þ

whereY0ðH; SÞis analytically derived by using the toroidal

approxima-tion in the following manner.

For 0∘≤θ≤40∘, it is observed from the numerical solutions of the

Young-Laplace equation that 0bK4=H≤1 and −4≤K3=Hb−2. Since

the contact radius is small, Yc≪1, the quartic term in Ycin Eq.(24)

can be neglected in comparison to the cubic term. Accordingly, Eq.(24)can be reduced to

K3Y3c þ K2Yc2þ K1Ycþ K0¼ 0 : ð27Þ

The discriminant of Eq.(27)was found to be negative by numerical verification. Thus, Eq.(27)has three unequal real solutions (see also [50]). These solutions can be determined by the trigonometric formula [50], of which the physically-relevant one is given by

Yc¼ − 2 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi −3K1K3þ K22 q 3K3 cosφ−K2 3K3 where cos3φ ¼27K0K23−9K1K2K3þ 2K32 2 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi − 3K1K3−K22  3 r ; ð28Þ

and K3, K2, K1and K0are given in Eq.(34). The accuracy of this solution

for Yc has been verified by comparisons of numerical solutions of

Eq.(24)and numerical solutions of the Young-Laplace equation. Substi-tution of this solution for Ycinto Eq.(22)gives an expression for the

di-mensionless neck radius Y0as a function of Hand S, i.e. Y0ðH; SÞ.

Given the dimensionless suction Hand the dimensionless separa-tion distance S, the procedure for determining the capillary force be-tween equal-sized spherical particles using the toroidal approximation consists of: (1) checking that the separation distance S is smaller than the critical separation Scritdetermined by Eq.(25); (2) calculating

the dimensionless contact radius Ycthrough Eq.(28)from the given

di-mensionless suction H and dimensionless separation distance S; (3) substituting Ycand Hinto Eq.(22)to obtain the dimensionless neck radius Y0; (4) with the obtained value of Y0, determining λ

from Eq.(26)and the capillary force Fcapfrom Eq.(11).

Fig. 4shows the comparison of the capillary forces from the numer-ical solutions of the Young-Laplace equation and those predicted by the toroidal approximation. The analytical theory agrees very well with the numerical solutions of the Young-Laplace equation when the mean cur-vature H≥104

and the contact angle 0∘≤θ≤40∘.

4. Curve-fitting expressions for rupture distances and capillary forces

The toroidal approximation described in Section3is limited to the dimensionless suction H≥104

. In this Section, computationally-fast closed-form expressions for the rupture distance and the capillary force of capillary bridges between equal-sized spherical particles under suction control have been obtained by curve-fitting to the large dataset of numerical solutions of the Young-Laplace equation. These expres-sions are applicable over a wider range of dimensionless suction than the analytical expressions that have been formulated in Section3. 4.1. Rupture distance

Gras et al. [30] obtained a closed-form expression for the rupture distance 2Scritfor given values of dimensionless suction H, by

curve-fitting to their limited dataset of numerical solutions of the Young-Laplace equation. As described in Ref. [30], such an expression is applicable only for the dimensionless suction in the range of 0≤H≤1

5. This range of dimensionless suction corresponds to large capillary bridge volumes V1=3≥0:1 and is very limited, compared to the whole range of dimensionless suction indicated inFig. 5. Therefore, it is neces-sary to formulate a closed-form expression that is capable of predicting rupture distances for a wider range of dimensionless suction than ex-pression 25ð Þ and that in Ref. [30].

By curve-fitting to a large dataset of numerical solutions of the Young-Laplace equation that is obtained by a high-resolution integra-tion method in Secintegra-tion2, a more accurate closed-form expression for determining the rupture distance 2Scrithas been obtained, given by 2Scrit¼ 2 H 1 ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 p  1 þ c1H−1=3þ c2H−2=3   ; ð29Þ

Fig. 3. Product of rupture distance and dimensionless suction, 2ScritH, for various values of the dimensionless suction H: numerical solutions of the Young-Laplace equation and predictions by the analytical theory; contact anglesθ ¼ 0∘; 20and 40.

(7)

where the coefficients c1and c2are dependent upon the contact angleθ,

given in Appendix B.

As shown inFig. 5, the obtained closed-form expression Eq.(29)

agrees very well with the numerical solutions of the Young-Laplace equation for the very large range of dimensionless suction considered with H≥1. For H≥102

, the deviations between thefit and the numer-ical solutions are smaller than 0:5%, much better than that of Gras et al. [30] which gives deviations up to 20%. For smaller values of the dimen-sionless suction, the deviations are less than 2:5%.

4.2. Capillary force

The analytical expression for the capillary force in Eq.(26)is not ac-curate for 1≤H≤104

(data not shown). Therefore, it is desirable to for-mulate a closed-form expression for the capillary force for a wider range of dimensionless suction, H≥1. Note that the dimensionless capillary forceλis determined by the dimensionless neck radius Y

0.

Based on observations on numerical solutions of the Young-Laplace equation, the neck radius Y0for capillary bridges between equal-sized

spherical particles is dependent on the separation distance S, the di-mensionless suction Hand the contact angleθ, i.e. Y

0¼ Y0ðS; H; θÞ.

Given the number of dependent variables, the steps for curve-fitting to the neck radius from numerical solutions of the Young-Laplace equa-tion are: (1) curve-fitting the neck radius at zero separation distance Y0fS¼0gfor dimensionless suction H≥1 and for contact angles θ of 0∘,

20∘and 40∘, i.e. Y0fS¼0g¼ Y0ðS¼ 0; H; θÞ; (2) curve-fitting the neck

ra-dius normalized by its value at zero separation distance Y0=Y0fS¼0gfor

dimensionless suction H≥1 and for contact angles θ of0∘, 20∘and 40∘, i.e. Y0=Y0fS¼0g¼ Y  0 Y0fS¼0g S =S crit; H; θ  

; and (3) obtaining the neck radius Y0by multiplying Y0=Y0fS ¼0gand Y0fS¼0g.

By curve-fitting to the large dataset extracted from numerical solu-tions of the Young-Laplace equation, an expression for the neck radius at zero separation distance, Y0fS¼0g, has been obtained

Y0 Sf¼0g¼ ffiffiffiffiffiffi 2 H r  p1exp − pffiffiffiffiffiffi2 H p  þ p3 ;  ð30Þ

Here the coefficients p1, p2and p3are functions of T (T¼ tan θ), given in

Appendix C. The accuracy of this expression is demonstrated inFig. 6. By curve-fitting to the large dataset, the neck radius normalized by the neck radius at zero separation distance Y0=Y0fS¼0gfor the selected

dimensionless suction H≥1 and for contact angles θ of 0∘, 20and 40,

the ratio of Y0=Y0fS¼0gis obtained as

Y0 Y0 Sf¼0g ¼ 1 þ A " ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1−S  Scrit s −1 # þ BS  Scrit; ð31Þ Fig. 4. Dependence of the capillary force on the separation distance, obtained from the numerical solutions of the Young-Laplace equation and predictions by the toroidal approximation, for contact angleθ of 0∘and 40. (a) Dimensionless suction H¼ 104

(b) dimensionless suction H¼ 105 .

Fig. 5. Rupture distances multiplied by dimensionless suction,2ScritH 

, for various values of dimensionless suction Hof numerical solutions to the Young-Laplace equation and the predictions by Eq.(29).

(8)

where A and B are functions of Handθ, given by A ¼ a1ln2ð Þ þ aH 2ln Hð Þ þ a 3 B ¼ b1ln 2 H ð Þ þ b2ln Hð Þ þ b 3 : ð32Þ

Here the coefficients aiand bionly depend on the contact angleθ, as

given in Appendix C. The accuracy of this expression is demonstrated inFig. 7.

Fig. 8shows the comparison for the dimensionless capillary forcesλ

between those from numerical solutions to the Young-Laplace equation

and the predictions by Eqs.(26), (30) and (31), for the dimensionless suction Hof 1, 25, 103and 105and the contact angleθ of 0∘and 40∘. The deviations between the predictions and the numerical solutions are smaller than 1% for bothθ ¼ 0∘andθ ¼ 40.

With these curve-fitting expressions, the steps for determining the capillary force of capillary bridges between equal-sized spherical parti-cles for given dimensionless suction H≥1 and separation distance S

are:

• checking that the separation distance Sis smaller than the critical

separation Scritdetermined by Eq.(29);

• calculating the neck radius at zero separation distance Y 0fS¼0g by

Eq.(30)from the given dimensionless suction H; • substituting Y

0fS¼0ginto Eq.(31)to obtain Y0;

• with the obtained values of Y0and specified H

, determining

λfrom Eq.(26)and the capillary force Fcapfrom Eq.(11).

5. Capillary bridges between unequal-sized spherical particles under suction control

In this Section, the expressions for capillary force and rupture dis-tance for the case of equal-sized spherical particles under suction con-trol are developed to the case of unequal-sized spherical particles. 5.1. Considering size effect by derjaguin radius

The effect of size ratio on capillary force and rupture distance has been studied in Refs. [17,18,28,51] (under volume control). It has been analytically and numerically proven in Ref. [18] that, for small capillary bridge volumes, the expressions for capillary force and rupture distance for equal-sized particles can be employed for unequal-sized particles by using the Derjaguin radius (also called the harmonic radius) that is defined in Eq.(2). The accuracy of the use of the Derjaguin radius R has been demonstrated for small capil-lary bridge volumes (V1=3≤0:1). Such a volume range corresponds to the dimensionless suction H N

≈ 10

2.

Fig. 7. Dependence on normalized separation distance S=S

critof the neck radius Y0normalized by that at zero separation distance, Y0fS¼0g, extracted from numerical solutions of

the Young-Laplace equation and that predicted by the closed-form expression in Eq.(31), for dimensionless suctions H¼ 1, 25, 103and 105: (a) contact angleθ ¼ 0, (b) contact angleθ ¼ 40∘.

Fig. 6. Product of Hð =2Þ1=2and the scaled neck radius at zero separation distance, Y 0fS¼0g, extracted from numerical solutions of the Young-Laplace equation and predicted by the closed-form expression in Eq.(30), for contact angles ofθ ¼ 0∘,θ ¼ 20andθ ¼ 40.

(9)

Using the Derjaguin radius R, the separation distance 2S, the capillary bridge volume V, the capillary forceλ and the mean curvature H can be made dimensionless 2S¼2S RV ¼ V R3λ ¼ λ RH ¼ H  R‾ : ð33Þ

Substituting the dimensionless S,λand Hinto the expressions for

rup-ture distances (Eqs.(25) and (29)) and capillary forces (Eqs.(26), (30) and (31)) for equal-sized spherical particles, rupture distances and cap-illary forces for unequal-sized spherical particles can be determined. The

accuracy of these predictions is investigated in the following subsections.

5.2. Rupture distance

Fig. 9a shows the comparison of rupture distances 2Scritextracted

from numerical solutions to the Young-Laplace equation and the predic-tions according to the toroidal approximation, for particle size ratios of RL=Rs = 1, 2, 8 and 128, contact angle of 40∘and dimensionless suction

H≥104. As expected, the toroidal approximation agrees very well with

numerical solutions to the Young-Laplace equation, with deviations smaller than 1%.

Fig. 9. Product of rupture distance and dimensionless suction, 2ScritH, as extracted from numerical solutions to the Young-Laplace equation and that predicted by the closed-form expressions for particle size ratio RL=Rs = 1, 2, 8 and 128 and contact angleθ ¼ 40∘: (a) toroidal approximation for dimensionless suction H≥104; (b) predictions by Eq.(29)for dimensionless suction H≥1.

Fig. 8. Capillary forces extracted from numerical solutions of the Young-Laplace equation and those predicted by the closed-form expressions Eqs.(26), (30) and (31), for dimensionless suctions H¼ 1, 25, 103and 105: (a) contact angleθ ¼ 0, (b) contact angleθ ¼ 40.

(10)

Fig. 9b compares the rupture distance 2Scritextracted from

nu-merical solutions of the Young-Laplace equation with the predic-tions of the closed-form expression in Eq.(29), for particle size ratios of RL=Rs = 1, 2, 8 and 128 and contact angle of 40∘. The

range of dimensionless suction is much larger than that for the to-roidal approximation in Fig. 9a. The closed-form expression agrees well with the numerical solutions to the Young-Laplace equation, with deviations smaller than 1% for H≥102

. For 1≤H≤102

(corre-sponding to capillary bridge volume V1=3≥0:1 ), the use of the Derjaguin radius gives larger deviations [18]. Hence the deviations between the numerical solutions of the Young-Laplace equation and the predictions of Eq.(29)are larger.

5.3. Capillary force

Fig. 10a shows the comparison between capillary forcesλextracted

from numerical solutions to the Young-Laplace equation and the predic-tions of the toroidal approximation, for particle size ratios of RL=Rs = 1,

2, 8 and 128, contact angles ofθ = 0∘and 40when the dimensionless

suction H≅105

. As expected, the toroidal approximation agrees very well with numerical solutions to the Young-Laplace equation, with de-viations smaller than 1%.

Fig. 10b compares the capillary forceλextracted from numerical

so-lutions to the Young-Laplace equation with the predictions of the closed-form, curvefitting expression in Eqs.(26), (30) and (31), for

Fig. 11. Numerical solutions to the Young-Laplace equation under volume control, with capillary bridge volume V≅10−6, particle size ratio R

L=Rs¼ 1 and contact angle θ ¼ 0∘: (a) dimensionless capillary forceλas a function of dimensionless separation distance S; note that two branches of solutions are present: the branch where the capillary forces is larger than the capillary force at critical separation is indicated by black dots, while the branch where the capillary forces smaller than the capillary force at critical separation is indicated by red triangles; (b) dimensionless free energy E¼ Esurf=γR‾2as a function of dimensionless separation distance S; (c) dimensionless free energy EsurfþΔpV¼ EsurfþΔpV=γR‾2 as a function of dimensionless separation distance S.

Fig. 10. Capillary forces from numerical solutions to the Young-Laplace equation and according to the predictions by (a) the toroidal approximation of Eq.(26)when the dimensionless suction of H≅105and (b) curve-fitting expression of Eqs.(26), (30) and (31)when the dimensionless suction of H≅102, for particle size ratios of R

L=Rs= 1, 2, 8 and 128, contact angles of θ ¼ 0∘andθ ¼ 40.

(11)

particle size ratios ofRL=Rs= 1, 2, 8 and 128 and contact angles ofθ = 0∘

and40∘when the dimensionless suction H≅102. It is shown that the

pre-dictions are in good agreement with the numerical solutions of the Young-Laplace equation.

From these analyses, it is concluded that the closed-form expres-sions are effective for capillary bridges between unequal-sized particles by using the Derjaguin radius, when the dimensionless suction H≥102

(corresponding to V1=3≤0:1, and thus the Derjaguin radius is valid [18]). The procedure for determining rupture distance and capillary force of capillary bridges between unequal-sized particles is then the same as that for equal-sized particles described in Sections 3 and 4. 6. Conclusions

This study focuses on capillary forces and rupture distances for cap-illary bridges between spherical particles under suction control, using three complementary different approaches: (1) analyses of numerical solutions of the governing Young-Laplace equation using a high-resolution integration method, (2) an analytical theory based on the to-roidal approximation, that does not require any calibrated parameters and (3) closed-form expressions with coefficients determined by curvefitting to the large dataset of numerical solutions of the Young-Laplace equation. Specific contributions of this study are:

• the effectiveness of the surface free-energy criterion has been demon-strated, under suction control and volume control, for identifying the physically-relevant solutions to the governing Young-Laplace equa-tion.

• based on the toroidal approximation, an analytical theory has been developed for the rupture distance and the capillary force (given values of the suction), that does not involve any calibrated parame-ters. The analytical expressions are valid when the dimensionless suc-tion H≥104

and the contact angleθ≤40∘.

• closed-form expressions have been formulated for the rupture dis-tance and the capillary force (given values of the suction) by curve-fitting to the large dataset of numerical solutions of the Young-Laplace equation. These expressions are in good agreement with the numerical dataset when the dimensionless suction H≥1 and the con-tact angleθ≤40∘.

• the influence of particle size ratio on the rupture distance and the cap-illary force has been considered by using the Derjaguin radius that is

valid when the dimensionless suction H≥102 (corresponding to

small capillary bridge volumes V1=3≤0:1).

The numerical solutions to the Young-Laplace equation and the closed-form expressions for capillary forces and rupture distances have not been compared with the experimental results in Ref. [30] for capillary bridges between spherical particles under suction control, since the values of the dimensionless suction Hin Ref. [30] are very small (in the range of 0≤H≤15) and the contact angle θ is not constant

in these experiments when the separation between the spherical parti-cles is varied.

For future studies, it is recommended to implement the developed closed-form expressions into Discrete Element Method simulation soft-ware or into micromechanics-based constitutive relations (see for in-stance Refs. [52–54]) for granular materials in order to investigate the effect of capillary cohesion under suction control. In addition, consider-ing the different behaviour of capillary bridges under volume and suc-tion control, the mechanical response of wet granular assemblies is highly loading-path dependent [55]. Therefore, it is important to con-sidering such properties in effective stress tensors [56–58] and constitu-tive relations [52,53,59] for wet granular materials.

Acknowledgements

The authors acknowledge support from the international research network GDRI-GeoMech (Multi-Physics and Multi-Scale Couplings in Geo-Environmental Mechanics). The research described was funded in part by the European Union's Horizon 2020 research and innovation program under the Marie Skłodowska-Curie grant agreement No. 832405 (ICARUS).

Appendix A. stable branches of numerical solutions to the Young-Laplace equation

As discussed in Section2.4, branches of (multiple) solutions of the Young-Laplace equation may exist under volume as well as suction con-trol. To identify the physically-relevant stable branch of solutions, criteria have been proposed based on the minimum of the energy in Eq.(13)for volume control and on the minimum of the energy in Eq.(14)for suction control. In this Appendix the appropriateness is in-vestigated of (“opposite”) criteria based on the energy in Eq.(13)for Fig. 12. Numerical solutions to the Young-Laplace equation under suction control, with dimensionless suction H≅103

, particle size ratio RL=Rs¼ 1 and contact angle θ ¼ 0∘: (a) dimensionless capillary forceλas a function of dimensionless separation distance S; note that two branches of solutions are present: the branch where the capillary forces is larger than the capillary force at critical separation is indicated by black dots, while the branch where the capillary forces smaller than the capillary force at critical separation is

indicated by red triangles; (b) dimensionless free energy E¼ Esurf=γR ‾ 2

as a function of dimensionless separation distance S; (c) dimensionless free energy EsurfþΔpV¼ EsurfþΔpV=γR‾ 2 as a function of dimensionless separation distance S.

(12)

suction control and on the energy in Eq.(14)for volume control. To this end, the large dataset of numerical solutions to the Young-Laplace equa-tion (see Secequa-tion2.3) is employed.

Under volume control (so withfixed capillary bridge volume V), the dependence of the capillary force on the separation S is shown in

Fig. 11a. Two branches of solutions exist: the capillary forces larger than the capillary force at critical separation are shown by (many) black dots, whereas the capillary forces smaller than the capillary force at critical separation are indicated by red triangles. The energies Esurfand EsurfþΔpVin Eqs.(13) and (14)are shown inFig. 11b and c,

re-spectively. According to Refs. [16,47,48], the stable branch under vol-ume control corresponds to a minimum in the energy Esurfin Eq.(13).

For this stable branch, the energy EsurfþΔpVin Eq.(14)is larger than

for the unstable branch. Therefore, the criterion of a minimum of the en-ergy EsurfþΔpVin Eq.(14)(proposed in Ref. [22] for suction control) can

not be used to identify the stable branch of solutions under volume control.

Under suction control (so withfixed suction H), the dependence of the capillary force on the separation S is shown inFig. 12a. The two branches of solutions exist: the capillary forces larger than the capillary force at critical separation are shown by (many) black dots, whereas the capillary forces smaller than the capillary force at critical separation are indicated by red triangles. The energies Esurf and EsurfþΔpVin Eqs.(13)

and (14)are shown in Fig. 12b and c, respectively. According to Ref. [22], the stable branch under suction control corresponds to a min-imum in the energy EsurfþΔpVin Eq.(14). For this stable branch, the

en-ergy Esurfin Eq.(14)is smaller than for the unstable branch. Therefore,

the criterion of a minimum of the energy Esurfin Eq.(13)(proposed in

Ref. [22] for volume control) can also be used to identify the stable branch of solutions under volume control. Note that under suction con-trol the stable branch has larger volume and surface area than the unsta-ble branch of solutions [22].

Similar analyses have been performed (data not shown) for other values of the contact angleθ, particle size ratio RL=Rs, capillary bridge

volume V and suction H, yielding the same conclusions.

From these analyses, it is concluded that the criterion of a minimum of the surface free energy Esurfdefined in Eq.(13)is appropriate to

iden-tify the stable branch of the solutions to the Young-Laplace equation, under both volume and suction control conditions.

Appendix B. rupture distance for given suction value H

In Eq.(24)for the dimensionless contact radius Yc, the coefficients

K4, K3, K2, K1and K0are dependent upon dimensionless suction H,

dimensionless separation distance S and contact angleθ (via T de-fined in Eq.(7)). These coefficients are given by

K4¼ H 2 ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 q T2−2T3þ ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 q −2T   K3¼ −4H T2− ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 q Tþ 1   K2¼ 8HST2þ 4HSþ 4T   ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 q −8HST3 −8HST−4T2 −12 K1¼ 8Hð STþ 8Þ ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 q þ −8T 2−8SH−8T K0¼ −4S½ −2HT2−H   S−2T   ffiffiffiffiffiffiffiffiffiffiffiffiffiffi T2þ 1 q þ 2 T 2þ 1 HSTþ 1 ð Þ : ð34Þ The quartic equation 24ð Þ has four real solutions, of which two are (potentially) physically relevant. At the critical separation distance Scrit, there is only a single solution, and hence the discriminantΔ of

the quartic equation 24ð Þ (see for example [50]) becomes zero. For values of the separation distance larger than the critical separation, no real solution with physical meaning exists.

Upon expansion of the equationΔ ¼ 0in the small parameter1=H, it

follows after some algebra (performed with the symbolic mathematics program Maple) that the discriminant is zero (up to second order in the small parameter 1=H) if

Scrit¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiα T2þ 1 p 1 H α ¼ 1−½ β Tð Þ H  1 3; ð35Þ

whereβ is a (complicated) function of T, whose third-order Taylor ex-pansion inβ is given by β ¼2732þ27 16Tþ 81 64T 2 : ð36Þ

In the expression for the rupture distance obtained from curve-fitting to the large dataset of numerical solutions of the Young-Laplace equation, Eq.(29), the coefficients cionly depend on the contact angle

θ. The resulting expressions for these coefficients are given by c1 ¼ −0:236T2−0:120T−1:358

c2 ¼ 0:239T2þ 0:111T þ 0:458 ;

ð37Þ in which the calibrated coefficients have been incorporated; T ¼ tan θ, whereθ is the contact angle in radians.

Appendix C. constants in the closed-form expression for neck radius In the expression for the neck radius at zero separation distance Y0fS¼0gobtained from curve-fitting to the large dataset of numerical

so-lutions of the Young-Laplace equation, Eq.(30), the coefficients pionly

depend on the contact angleθ. The resulting expressions for the func-tions pi (expressed concisely as a matrix-vector multiplication) are

given by p1 p2 p3 2 4 3 5 ¼ −0:151 −0:015 0:8500:302 0:038 1:217 0:012 −0:017 0:152 2 4 3 5 T 2 T 1 2 4 3 5; ð38Þ

where the calibrated parameters are incorporated, and T¼ tan θ as de-fined in Eq.(7).

In the same manner, in the expression for the ratio of neck radius to its value at zero separation distance Y0=Y0fS¼0g obtained by

curve-fitting to the large dataset of numerical solutions of the Young-Laplace equation, Eq.(31), the coefficients aiand bipresent in Eq.(32)only

de-pend on the contact angleθ (also through T ¼ tan θ). The resulting ex-pressions for the functions ai(expressed concisely as a matrix-vector

multiplication) are given by a1 a2 a3 2 4 3 5 ¼ −0:00150:0347 −0:0142 0:00210:0001 0:0019 −0:1165 0:0311 0:6087 2 4 3 5 T 2 T 1 2 4 3 5 ; ð39Þ

and for the functions bi(in terms of T) are

b1 b2 b3 2 4 3 5 ¼ −0:00160:0330 −0:0201 −0:03650:0006 0:0026 −0:0949 0:0674 0:0088 2 4 3 5 T 2 T 1 2 4 3 5 : ð40Þ References

[1] H.C. Hamaker, The London–van der Waals attraction between spherical particles, Physica 4 (10) (1937) 1058–1072.

[2] I. Aranson, D. Blair, V. Kalatsky, G. Crabtree, W.-K. Kwok, V. Vinokur, U. Welp, Elec-trostatically driven granular media: phase transitions and coarsening, Phys. Rev. Lett. 84 (15) (2000) 3306–3309.

[3] R. Fisher, On the capillary forces in an ideal soil; correction of formulae given by W.B. Haines, J. Agric. Sci. 16 (3) (1926) 492–505.

(13)

[4] F. Soulié, F. Cherblanc, M. El Youssoufi, C. Saix, Influence of liquid bridges on the me-chanical behaviour of polydisperse granular materials, Int. J. Numer. Anal. Methods Geomech. 30 (3) (2006) 213–228.

[5] V. Richefeu, M. El Youssoufi, E. Azéma, F. Radjaï, Force transmission in dry and wet granular media, Powder Technol. 190 (1–2) (2009) 258–263.

[6] L. Scholtès, B. Chareyre, F. Nicot, F. Darve, Micromechanics of granular materials with capillary effects, Int. J. Eng. Sci. 47 (1) (2009) 64–75.

[7] J.-Y. Delenne, V. Richefeu, F. Radjaï, Liquid clustering and capillary pressure in gran-ular media, J. Fluid Mech. 762 (R5) (2015) 1–10.

[8] M. Pouragha, R. Wan, J. Duriez, N.H. Sultan, Statistical analysis of stress transmission in wet granular materials, Int. J. Numer. Anal. Methods Geomech. 42 (16) (2018) 1935–1956.

[9] M.E.D. Urso, C.J. Lawrence, M.J. Adams, Pendular, funicular, and capillary bridges: re-sults for two dimensions, J. Colloid Interface Sci. 220 (1) (1999) 42–56.

[10] J.-P. Wang, E. Gallo, B. François, F. Gabrieli, P. Lambert, Capillary force and rupture of funicular liquid bridges between three spherical bodies, Powder Technol. 305 (2017) 89–98.

[11] K. Murase, T. Mochida, Y. Sagawa, H. Sugama, Estimation on the strength of a liquid bridge adhered to three spheres, Adv. Powder Technol. 19 (2008) 349–367. [12] D. Lievano, S. Velankar, J.J. McCarthy, The rupture force of liquid bridges in two and

three particle systems, Powder Technol. 313 (2017) 18–26.

[13]M.J. Adams, S.A. Johnson, J.P. Seville, C.D. Willett, Mapping the influence of gravity on pendular liquid bridges between rigid spheres, Langmuir 18 (16) (2002) 6180–6184.

[14]N.P. Kruyt, O. Millet, An analytical theory for the capillary bridge force between spheres, J. Fluid Mech. 812 (2017) 129–151.

[15] C.-F. Zhao, N.P. Kruyt, O. Millet, Capillary bridge force between non-perfectly wetta-ble spherical particles: an analytical theory for the pendular regime, Powder Technol. 339 (2018) 827–837.

[16] G. Lian, C. Thornton, M.J. Adams, A theoretical study of the liquid bridge forces be-tween two rigid spherical bodies, J. Colloid Interface Sci. 161 (1) (1993) 138–147. [17] G. Lian, J. Seville, The capillary bridge between two spheres: new closed-form

equa-tions in a two century old problem, Adv. Colloid Interface Sci. 227 (2016) 53–62. [18]C.-F. Zhao, N.P. Kruyt, O. Millet, Capillary bridges between unequal-sized spherical

particles: rupture distances and capillary forces, Powder Technol. 346 (2019) 462–476.

[19] P.A. Cundall, O.D.L. Strack, A discrete numerical model for granular assemblies, Géotechnique 29 (1) (1979) 47–65.

[20] L. Scholtès, P.-Y. Hicher, F. Nicot, B. Chareyre, F. Darve, On the capillary stress tensor in wet granular materials, Int. J. Numer. Anal. Methods Geomech. 33 (10) (2009) 1289–1313.

[21] J.-P. Wang, X. Li, H.-S. Yu, A micro–macro investigation of the capillary strengthen-ing effect in wet granular materials, Acta Geotechnica 13 (3) (2018) 513–533. [22]J. Duriez, R. Wan, Contact angle mechanical influence in wet granular soils, Acta

Geotechnica 12 (1) (2017) 67–83.

[23]J. Duriez, R. Wan, Subtleties in discrete-element modelling of wet granular soils, Géotechnique 67 (4) (2017) 365–370.

[24]C. Delaunay, Sur la surface de révolution dont la courbure moyenne est constante, J. Math. Pures Appl. (1841) 309–314.

[25] M.A. Erle, D. Dyson, N.R. Morrow, Liquid bridges between cylinders, in a torus, and between spheres, AIChE J. 17 (1) (1971) 115–121.

[26] F. Orr, L. Scriven, A.P. Rivas, Pendular rings between solids: meniscus properties and capillary force, J. Fluid Mech. 67 (4) (1975) 723–742.

[27] G. Mason, W. Clark, Liquid bridges between spheres, Chem. Eng. Sci. 20 (10) (1965) 859–866.

[28]C.D. Willett, M.J. Adams, S.A. Johnson, J.P. Seville, Capillary bridges between two spherical bodies, Langmuir 16 (24) (2000) 9396–9405.

[29] Y.I. Rabinovich, M.S. Esayanur, B.M. Moudgil, Capillary forces between two spheres with afixed volume liquid bridge: theory and experiment, Langmuir 21 (24) (2005) 10992–10997.

[30] J.-P. Gras, J.-Y. Delenne, M.S. El Youssoufi, Study of capillary interaction between two grains: a new experimental device with suction control, Granul. Matter 15 (1) (2013) 49–56.

[31] H.N.G. Nguyen, O. Millet, G. Gagneux, On the capillary bridge between spherical par-ticles of unequal size: analytical and experimental approaches, Continuum Mech. Therm. (2018) 1–13.

[32] H.N.G. Nguyen, O. Millet, G. Gagneux, Exact Calculation of Axisymmetric Capillary Bridge Properties between Two Unequal-Sized Spherical Particles, Mathematics and Mechanics of Solids, 2018https://doi.org/10.1177/1081286518787842. [33] G. Gagneux, O. Millet, An analytical framework for evaluating the cohesion effects of

coalescence between capillary bridges, Granul. Matter 18 (2) (2016) 16. [34] B. Mielniczuk, O. Millet, G. Gagneux, M.S. El Youssoufi, Characterisation of pendular

capillary bridges derived from experimental data using inverse problem method, Granul. Matter 20 (1) (2018) 14.

[35] G. Gagneux, O. Millet, B. Mielniczuk, M.S. El Youssoufi, Theoretical and experimental study of pendular regime in unsaturated granular media, European Journal of Envi-ronmental and Civil Engineering 21 (7–8) (2017) 840–853.

[36]B.V. Derjaguin, Untersuchungen über die Reibung und Adhäsion, IV (in German), Colloid Polym. Sci. 69 (2) (1934) 155–164.

[37]F. Molenkamp, A. Nazemi, Interactions between two rough spheres, water bridge and water vapour, Géotechnique 53 (2) (2003) 255–264.

[38] A. Nazemi, A. Majnooni-Heris, A mathematical model for the interactions between non-identical rough spheres, liquid bridge and liquid vapor, J. Colloid Interface Sci. 369 (1) (2012) 402–410.

[39] J.-P. Gras, J.-Y. Delenne, F. Soulié, M.S. El Youssoufi, DEM and experimental analysis of the water retention curve in polydisperse granular media, Powder Technol. 208 (2) (2011) 296–300.

[40]J. Lechman, N. Lu, Capillary force and water retention between two uneven-sized particles, J. Eng. Mech. 134 (5) (2008) 374–384.

[41] O. Harireche, A. Faramarzi, A.M. Alani, A toroidal approximation of capillary forces in polydisperse granular assemblies, Granul. Matter 15 (5) (2013) 573–581. [42] F.R. De Bisschop, W.J. Rigole, A physical model for liquid capillary bridges between

adsorptive solid spheres: the nodoid of Plateau, J. Colloid Interface Sci. 88 (1) (1982) 117–128.

[43] N. Lu, W.J. Likos, Unsaturated Soil Mechanics, Wiley, 2004.

[44] D.G. Fredlund, H. Rahardjo, M.D. Fredlund, Unsaturated Soil Mechanics in Engineer-ing Practice, John Wiley & Sons, 2012.

[45] J.A. Blatz, Y.-J. Cui, L. Oldecop, Vapour equilibrium and osmotic technique for suction control, Laboratory and Field Testing of Unsaturated Soils, Springer 2008, pp. 49–61. [46] M. Dörmann, H.-J. Schmid, Distance-dependency of capillary bridges in

thermody-namic equilibrium, Powder Technol. 312 (2017) 175–183.

[47] B.J. Lowry, P.H. Steen, Capillary surfaces: stability from families of equilibria with ap-plication to the liquid bridge, Proc. R. Soc. Lond. Ser A Math. Phys. Sci. 449 (1937) (1995) 411–439.

[48]J. Bostwick, P. Steen, Stability of constrained capillary surfaces, Annu. Rev. Fluid Mech. 47 (2015) 539–568.

[49]K. Galvin, A conceptually simple derivation of the Kelvin equation, Chem. Eng. Sci. 60 (2005) 4659–4660.

[50] G.A. Korn, T.M. Korn, Mathematical Handbook for Scientists and Engineers, McGraw-Hill, 2000.

[51]V. Mehrotra, K. Sastry, Pendular bond strength between unequal-sized spherical particles, Powder Technol. 25 (2) (1980) 203–214.

[52] P.-Y. Hicher, C.S. Chang, A microstructural elastoplastic model for unsaturated gran-ular materials, Int. J. Solids Struct. 44 (7–8) (2007) 2304–2323.

[53]C.-F. Zhao, Y. Salami, P.-Y. Hicher, Z.-Y. Yin, Multiscale Modeling of Unsaturated Granular Materials Based on Thermodynamic Principles, Continuum Mech. Therm. 31 (1) (2019) 341–359.

[54] C.-F. Zhao, Z.-Y. Yin, A. Misra, P.-Y. Hicher, Thermomechanical formulation for micromechanical elasto-plasticity in granular materials, Int. J. Solids Struct. 138 (2018) 64–75.

[55] H. Rahardjo, O.B. Heng, L.E. Choon, Shear strength of a compacted residual soil from consolidated drained and constant water content triaxial tests, Can. Geotech. J. 41 (3) (2004) 421–436.

[56] A.W. Bishop, G. Blight, Some aspects of effective stress in saturated and partly satu-rated soils, Géotechnique 13 (3) (1963) 177–197.

[57]E.E. Alonso, J.-M. Pereira, J. Vaunat, S. Olivella, A microstructurally based effective stress for unsaturated soils, Géotechnique 60 (12) (2010) 913–925.

[58] J. Duriez, R. Wan, Stress in wet granular media with interfaces via homogenization and discrete element approaches, J. Eng. Mech. 142 (12) (2016), 04016099. [59] A. Gens, M. Sánchez, D. Sheng, On constitutive modelling of unsaturated soils, Acta

Referenties

GERELATEERDE DOCUMENTEN

massaspectrometrie een gereedschap worden verkregen dat bijzonder geschikt is voor het speuren naar biologisch relevante onbekende contaminanten.. Een succesverhaal is de vondst

Die modellen zijn ook van belang voor het doen van onderzoek van atmosferische verschijnselen in relatie tot het weer, de luchtkwaliteit en het klimaat.. Daar- naast worden

Although all 3 thawed embryos were intact at transfer, blastomeres of embryos are often damaged after thawing.. There are, however, no indications that abnormalities increase

Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication:.. • A submitted manuscript is

Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of

Mobilization of (a) dissolved iron, (b) total inorganic carbon, (c) dissolved organic carbon, (d) ammonium, (e) methane and (f) total dissolved phosphorus over time (t = 0, 30 and

5.4 is a C-V curve (500 kc/s) of the gate electrode which curve indicates the state of the surface (depleted, or inverted) under the gate electrode. These results indicate

Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication:.. • A submitted manuscript is