Graphical Abstract Figure
Graphical Abstract Figure
Close modal

Abstract

Understanding peeling behavior in soft materials is integral to diverse applications, from tissue engineering, wound care, and drug delivery to electronics, automotive, and aerospace equipment. These applications often require either strong, permanent adhesion or moderate, temporary adhesion for ease of removal or transfer. Soft adhesives, especially when applied on soft substrates like elastomer-coated release liners, flexible packaging films, or human skin, present unique mechanical behaviors compared to adhesives applied on rigid substrates. This difference highlights the need to understand the influence of substrate rigidity on peeling mechanics. This review delves into both energy- and stress-based analyses, where a thin tape with an adhesive layer is modeled as a flexible beam. The energy analysis encompasses components like the energy associated with tape deformation, kinetic energy, and energy lost due to interfacial slippage. The stress analysis, on the other hand, focuses on structures with thin, deformable substrates. Substrates are categorized into two types: those undergoing smaller deformations, typical of thin soft release liners, and thicker deformable substrates experiencing significant deformations. For substrates with small deformations, the linear Euler–Bernoulli beam theory is applied to the tape in the bonded region. Conversely, for substrates experiencing significant deformations, large deflection theory is utilized. These theoretical approaches are then linked to several practical, industrially relevant applications. The discussion provides a strategic guide to selecting the appropriate peeling theory for a system, emphasizing its utility in comprehending peeling mechanisms and informing system design. The review concludes with prospective research avenues in this domain.

1 Introduction

Peeling mechanics play a crucial role in understanding the mechanisms underlying the separation of bonded surfaces. Separation, or “delamination,” events occur in everyday scenarios, such as the detachment of a sticker from a book or the removal of a medical bandage from skin. In an industrial context, the importance of peeling mechanics extends to the repair of fractured objects using adhesive tapes, detachment of labels from barrier films employed in packaging, and behavior of composite laminates utilized in the aerospace and automotive sectors. Promoting delamination is as crucial as suppressing it and both can be managed by controlling peel strength. A fundamental understanding of peeling mechanics is thus vital for the informed design and optimization of formulations with optimal peel behavior.

To comprehend the intricate phenomena of peeling, which can involve a multitude of materials, surfaces, and environments, various analytical models and finite element analysis methods have been employed. These approaches enable numerous parameters to be probed such as peeling angle [1,2], work of adhesion [1,3], peeling velocity [4,5], predeformation of the tape [6], friction [7,8], and film stiffness [9,10].

While peel performance depends on a range of material and geometric parameters directly associated with adhesive and interfacial interactions, the type of substrate to which the adhesive is bonded also plays a significant role (Fig. 1). Rigid metals and ceramics in the automotive, construction, and electronics industries usually exhibit high surface energies, facilitating superior wettability and adhesion properties, compared to the surface energy of engineering plastics, such as polyolefins and polyesters [11]. In contrast, semi-rigid engineering plastics, soft elastomers, and human skin—commonly encountered in packaging and labeling, healthcare, and surface coating applications—typically display lower surface energies on the order of 101 mJ/m2 [11,12], impeding the wettability and adhesion of applied adhesives. Furthermore, the differences in rigidity and geometry across the varying substrates significantly influence the adhesion performance. During the peeling process, the relative stiffness of the substrate compared to the adhesive results in distinct deformation profiles through the bulk of both the adhesive and the substrate. This interplay of mechanical response and energy dissipation plays a crucial role in determining the overall peel strength, which can vary by several orders of magnitude across substrates—typically ranging from approximately 100 J/m2 to 103 J/m2 for pressure-sensitive adhesives (PSAs). Therefore, understanding how a single adhesive interacts with different substrates is essential for ensuring its key functionalities, as numerous applications involve more than one surface. For example, a single adhesive must interact with a release liner to protect the adhesive surface before use and facilitate effortless release prior to use, but the same adhesive must also form a strong bond on another substrate during use.

Fig. 1
Peeling behaviors of adhesives from rigid, semi-rigid, and soft substrates. Observe the increasing substrate deformation as it transitions from rigid to soft. Here, the peeling arm typically corresponds to the soft adhesive mounted on a more rigid backing material.
Fig. 1
Peeling behaviors of adhesives from rigid, semi-rigid, and soft substrates. Observe the increasing substrate deformation as it transitions from rigid to soft. Here, the peeling arm typically corresponds to the soft adhesive mounted on a more rigid backing material.
Close modal

Several outstanding review papers touch upon analytical or computational (simulation) models related to adhesive and interfacial interactions. For example, Sun et al. [13] provided insights into analytical peeling models, highlighting their ability to accurately correlate peel force with the viscoelastic properties of the PSA. Park and Paulino [14], Needleman [15], and Wciślik and Pała [16] reviewed the cohesive zone model (CZM), a tool for characterizing interfacial properties in structures that experience adhesive failure. Several other review papers have also considered deformable substrates. Sauer [17], for example, provided a comprehensive overview of the computational models, especially those that are versatile, catering to diverse adhesion mechanisms and adherent geometries. Gu et al. [18] summarized the advances in analytical and computational peeling models for capturing adhesive properties on natural and artificial surfaces. The authors demonstrate, with the aid of the peeling model, that biological and biomimetic surface adhesion can be understood, and surface adhesive material can be designed by modulating the backing and substrate topography and adding interfacial linkers. More recently, Bartlett et al. [19] extensively reviewed different types of peel tests and discussed their applications in various materials, including biological systems and soft materials.

While numerous literature reviews exist, these reviews often underemphasize the specific peel mechanics associated with soft substrates. These structures are pivotal in a variety of applications, including release liners for pressure-sensitive adhesives, bio-adhesives for human skin, water-repellent coatings, and food packaging. The following review thus examines these contexts in depth, providing an overview of the theoretical methods and some industrial applications of peeling from soft substrates. In Sec. 2, we provide an overview of theoretical foundations to understand the adhesive system. Section 3 delves into the energy-based analysis of peeling, while Sec. 4 focuses on stress-based analysis for structures with a soft substrate. In Sec. 5, we show specific applications of the analytical models and conclude with remarks and future research prospects.

2 Basics of Peeling Theory

The process of peeling a tape from a soft substrate is shown in Fig. 2, where the width and thickness of the tape are denoted by b and h, respectively. The soft tape (pink) is peeled away from a soft substrate (blue) of thickness hl at an angle θ by a peel force F that induces a steady-state peel velocity V at the end of the peeling arm. The soft substrate is mounted on a rigid backing, denoted by diagonal hashed lines. Here, the peeling arm refers to the portion of tape that is not bonded to the soft substrate (see Fig. 2).

Fig. 2
An illustration of peeling from a soft substrate. The tape and soft substrate are represented in pink and blue colors, respectively. The thicknesses of the tape and the substrate are denoted as h and hl, respectively. The soft substrate is mounted on a rigid backing. The peeling arm signifies the section of the tape detaching from the soft substrate. Point A denotes the peel front, while point C marks the end of the peeling arm. Any point on the peeling arm with coordinates (x, y) can alternatively be represented by arc length s of the peeling arm with the corresponding tangential angle ϕ(s). Consequently, ϕ equates to θ0 and θ at point A and C, respectively.
Fig. 2
An illustration of peeling from a soft substrate. The tape and soft substrate are represented in pink and blue colors, respectively. The thicknesses of the tape and the substrate are denoted as h and hl, respectively. The soft substrate is mounted on a rigid backing. The peeling arm signifies the section of the tape detaching from the soft substrate. Point A denotes the peel front, while point C marks the end of the peeling arm. Any point on the peeling arm with coordinates (x, y) can alternatively be represented by arc length s of the peeling arm with the corresponding tangential angle ϕ(s). Consequently, ϕ equates to θ0 and θ at point A and C, respectively.
Close modal

For clarity, the tape structure shown in Fig. 2 can represent different scenarios depending on the application conditions. Failure may occur at the interface between the adhesive and the substrate material (adhesive failure) or within the adhesive material itself (cohesive failure). In instances of adhesive failure, the tape shown in Fig. 2 includes both the adhesive and backing materials. Conversely, in cases of cohesive failure, the tape shown in Fig. 2 represents the adhesive backing with part of the adhesive, whereas the substrate consists of the intended substrate and the remaining adhesive material.

The principle of energy conservation is fundamental to adhesive systems and is independent of the geometry of the system. This framework makes it a universally applicable approach for adhesive structures, regardless of their shape or size. In this section, the principle of energy conservation is employed for the adhesive system, drawing upon a similar approach outlined in the work by Kinloch et al. [20]. The energy conservation equation of the peeling system can be written as follows:
(1)
where dUext is the externally applied energy during the peeling process, and dUs=dUst+dUss is the change in the stored strain energy in the structure. dUst and dUss are the changes in the stored strain energy of the tape and soft substrate, respectively. At steady state, the strain energy stored in the soft substrate stabilizes, making the increment of stored strain energy (dUss) negligible. dUd is the change in energy dissipated within the bulk of the structure, excluding that at the interface. dUk is the change in the kinetic energy. dUsl is the increment of the energy dissipation from the interfacial sliding between the tape and substrate. dA is the increment of the area due to the creation of new surfaces during peeling. Γ is the interfacial fracture energy, which is a material fracture parameter used to characterize the energy required to create a new surface during peeling (see more details in Sec. 3.4).

Evaluating each term in Eq. (1) theoretically leads to a complete description of the peel mechanics of the adhesive system. Fracture mechanics theory can be used to evaluate the stress field around the peel front, thereby determining each term in Eq. (1). The stress field in the elastic layer structure near the peel front has a singularity, which is physically impossible because irreversible deformation, such as plastic deformation, will eliminate the singularity in real materials. To consider the local details near the peel front in materials with irreversible deformation, the CZM has been employed [2124]. However, here, the focus is on analytical models.

To understand delamination processes, two methods are commonly used: energy-based analysis and stress-based analysis. These methods differ in their failure criterion. In energy-based analyses, the peel front propagates once the external input energy is sufficient to create a new surface. On the other hand, in stress-based analysis, the peel front propagates when the stress or deformation at the peel front reaches a critical value. In this review, we will discuss these two methods in detail, with a focus on adhesive structures with a soft substrate, as shown in Fig. 2. This review covers the basics and recent advances of these two methods. Throughout this review, the soft substrate will be represented as a light blue material, while the pink color denotes the flexible tape.

3 Energy-Based Analysis

To derive analytical expressions for peel force, certain assumptions must be made. For example, the tape can be approximated as a flexible beam [5,10,20,25,26]. When the peel force attains a critical threshold, this beam-like tape detaches from the soft substrate, resulting in the formation of new surfaces. As the peel front (point A in Fig. 2) moves forward a distance da, a displacement of da(1+εacosθ) is induced at point C (see Fig. 2) of the peeling arm, where εa represents the tensile strain of the tape under the action of the load F at steady state. Therefore, the externally applied energy during the peeling process is dUext=Fda(1+εacosθ). Then Eq. (1) can be rewritten as follows:
(2)
where dUsta and dUstp are the changes in the strain energy from the tensile deformation of the attached (bonded arm) and the detached area (peeling arm) between the tape and the substrate. dUsba and dUsbp are the changes in the strain energy from the bending deformation of the bonded and peeling arm. At steady state, the increment of stored strain energy from the bending deformation in the tape d(Usba+Usbp) is negligible. dUdta and dUdtp are the increments of the dissipated energy from the tensile deformation of the bonded and peeling arms, respectively. dUdba and dUdbp are the increment of the dissipated energy from the bending deformation of bonded and peeling arms, respectively. dUds is the dissipation of energy from the deformation of the substrate. In the simplest case of an elastic adhesive structure with inextensible tape, based on the energy-based analysis, Rivlin [3] obtained the following:
(3)
Then, Kendall [27] considered the extensibility of the peeling arm and derived that
(4)
where the second term is from the extensional deformation of the peeling arm. In more complicated cases, more energy terms in Eq. (2) should be considered. In the following, the energy in the tape (including the peeling and bonded arm), kinetic energy (Uk), dissipation energy induced by interfacial sliding (Usl), fracture energy (Γ), and the energy in the soft substrate (Uds) will be discussed.

3.1 Energy Terms Associated With the Tape.

As the increment of stored strain energy from the bending deformation in the tape d(Usba+Usbp) is insignificant at steady state, our focus is on evaluating the following terms: d(Usta+Ustp+Udta+Udtp+Udba+Udbp). For certain special scenarios, such as peeling a thin elastic tape from a thin elastic substrate with a peeling angle close to zero (lap joint), Kendall [28] demonstrated that only d(Usta+Ustp) must be considered. Under the assumption of uniform tension deformation in the linear elastic (LE) substrate and tape, stress and deformation can be determined through force equilibrium equations, enabling calculation of d(Usta+Ustp). However, for many other cases, simplifying assumptions are necessary to derive analytical expressions for energy contributions from the tape. For example, some researchers neglect deformation in the bonded arm since the deformation in the bonded arm is significantly smaller than that in the peeling arm [2932]. In this case, only the energy term d(Ustp+Udtp+Udbp) needs to be evaluated.

In the following discussion, the energy terms associated with the peeling arm will be discussed, specifically Ustp+Udtp+Udbp. Let (x, y) denote the coordinate with the origin at the peel front (see Fig. 2). Any point on the peeling arm can alternatively be represented by the arc length s of the peeling arm with the corresponding tangential angle ϕ(s). At steady state, the peeling arm has a fixed configuration, so ϕ(s) is time independent.

For a given peel distance da, the following holds [33]:
(5)
where M(s) and κ(s)=dϕ/ds are the bending moment and curvature, respectively. σ(s)=Fcos(θϕ(s))/bh and ε(s) are the tensile stress and strain, respectively. To solve Eq. (5), the shape of the peeling arm ϕ(s) must be determined by solving the equilibrium equations of bending [33]:
(6)
As shown in Fig. 2, the tangential angle at the peel front (point A) and the end of the peeling arm (point C) are denoted as θ0 and θ, respectively. Therefore, the boundary conditions for Eq. (6) can be given as follows:
(7)

Typically, peeling tests are conducted at either θ=90deg or θ=180deg configuration. Provided that the material constitutive model of the tape is given, then the shape of the peeling arm can be determined through Eqs. (6) and (7), which enables the derivation of the stress σ(s) and strain ε(s), respectively. Subsequently, the value of Ustp+Udtp+Udbp in Eq. (5) can be evaluated. To illustrate this process, three examples of the peeling arm are presented: elastic, elastic–plastic, and viscoelastic peeling arm.

3.1.1 Elastic Peeling Arm.

For a linear elastic tape, the change in bending elastic energy does not contribute to the energy conservation equation [34,35]. Therefore, only dUstp in Eq. (5) needs to be evaluated:
(8)

3.1.2 Elastic–Plastic Peeling Arm.

When instead dealing with an elastic–plastic tape, both stretching and bending contribute to the overall peeling force. As a result, each term of Eq. (5) needs to be evaluated. The first two terms can be computed directly from the constitutive model as follows:
(9)

The evaluation of the third term, Udbp, which arises from the bending deformation, is a complex task due to the elastic–plastic constitutive relations that result in complicated mathematical forms—often making an explicit expression impossible. The current literature provides methods for assessing the dissipation resulting from the bending of the peeling arm. For instance, solutions for nonwork hardening elastic-perfect plastic [20], bilinear work hardening [20], power-law work hardening [36], and general work hardening elastic–plastic [37] peeling arms have been reported. Although the process of evaluating the dissipation from the bending of the peeling arm may be challenging, free spreadsheet tools [38] have been developed, which can be used to compute this term for the aforementioned cases.

3.1.3 Viscoelastic Peeling Arm.

For linear viscoelastic (LVE) tape, the deformation is time dependent, and the following constitutive relationship holds [33]:
(10)
where I is the moment of inertia, E(t) is relaxation modulus and
(11)
where E is the equilibrium modulus and Ei is the instantaneous modulus corresponding to the relaxation time τi. Equation (11) can be written in Prony series as follows:
(12)
where gi=Ei/E0 and E0 is the instantaneous modulus of the viscoelastic material.
By substituting Eq. (10) into Eq. (6), the following equation is obtained:
(13)
To evaluate the right-hand side of Eq. (5), ϕ(s) is needed in the place of ϕ(t). Therefore, we transform Eq. (13) in terms of s. The arc length s=Vt, where V is the velocity in the peeling arm and is assumed to be constant when the stretching deformation of the tape is small [33]. Consequently, we can rewrite Eq. (13) as follows:
(14)
Equation (14) can be solved with the boundary conditions in Eq. (7). Once ϕ(s) is evaluated, the second term of the right-hand side of Eq. (5) can be evaluated. Therefore, only the first term of the right-hand side of Eq. (5) remains to be evaluated. We will demonstrate this as follows. The stress σ(t)=Fcos(θϕ(t))/bh and the corresponding strain can be evaluated as follows:
(15)
where J(t) is creep compliance. Subsequently, the first term of the right-hand side of Eq. (5) can be evaluated.

Equations (10)(15) present a general framework for analyzing the energy contributions from a linear viscoelastic peeling arm. This framework was utilized by Chen et al. [33], who examined a viscoelastic peeling arm with one relaxation time based on an energy balance. The authors evaluated the energy dissipation arising from the bending and stretching of the peeling arm. Similarly, Ceglie et al. [8] also considered the energy dissipation resulting from a viscoelastic peeling arm with one relaxation time. However, their focus was on thin tapes, where the bending contribution can be neglected. In this case, only the first term on the right-hand side of Eq. (5) is considered.

3.2 The Kinetic Energy Term.

In this subsection, the kinetic energy term (dUk in Eq. (2)) associated with the peeling arm is discussed. To our knowledge, there are currently no analytical expressions in the literature that specifically address the kinetic energy contributions of the bonded arm and the soft substrate, as these contributions are typically small relative to other terms. This section reviews an approach for calculating the kinetic energy of the peeling arm, including the components of the peel velocity along and perpendicular to the peeling arm [39]:
(16)
Thus:
(17)
where ρ is the mass density of the tape. Consider the architecture in which all components are elastic. The energy consideration provides the following:
(18)
Considering a small deformation εa1, the following relationship holds:
(19)
where E is Young's modulus of the elastic tape and C=(E/ρ) is the elastic wave velocity of the tape. Vp=da/dt is the steady-state velocity of the propagation of the peel front. The dependence of the normalized peeling force F(1cosθ)/(Γ0b) on the normalized crack (peel front) propagation velocity Vp/(CΓ/(Eh(1cosθ))) is illustrated in Fig. 3. Clearly, for an elastic peeling system, the kinetic energy cannot be ignored when Vp is on the order of Γ/(Eh(1cosθ))C. When Vp is around 22% of the normalized crack propagation velocity, the peeling force will increase by approximately 5% compared to the quasi-static peeling (Vp is nearly zero).
Fig. 3
Dependence of normalized peel force on normalized crack (peel front) propagation velocity
Fig. 3
Dependence of normalized peel force on normalized crack (peel front) propagation velocity
Close modal

3.3 Energy Dissipation From Interfacial Frictional Stress.

In this subsection, the energy dissipation (dUsl in Eq. (2)) from interface slippage is discussed. Given the spatial distribution of the slip velocity v, the constitutive relations between shear stress τ and slip velocity v, and a peeling distance da at steady state, the increment of the energy dissipation from interfacial sliding between the tape and the substrate dUsl can be calculated as follows:
(20)
For unit peeling distance, the slip displacement u(x) can be evaluated as follows:
(21)
Then,
(22)
By substituting Eq. (22) into Eq. (20), the general equation to calculate the interfacial friction dissipation is obtained:
(23)

dUsl can be evaluated from Eq. (23) for given constitutive relations τ(v) and u(x)/x. Since τ(v) and u(x)/x can be treated differently, we will provide two specific examples.

Zhang Newby and Chaudhury [40] considered slippage of a viscoelastic adhesive on a rigid substrate, which can be adapted to analyze structures involving a soft substrate. The authors measured the displacement distribution u(x) by probing the lateral motion of fluorescent particles in the adhesive. In addition, they applied pure shear and obtained the constitutive relations τ(v) as follows:
(24)
where k1 and k2 are fitting constants from pure shear tests. Then they obtain the interfacial energy dissipation term:
(25)
where n is 0.35 when v<100μm/s, or 0.52 when v>100μm/s.
Ceglie et al. [8] investigated the slippage of viscoelastic tape on a rigid substrate; their approach can be extended to the case of a soft substrate. The authors assume that the interfacial shear stress τ is uniformly distributed in a local area α<x<0 (see Fig. 4). Then, Eq. (23) can be rewritten as follows:
(26)
where τ=Fcosθ/(αb)=σ0hcosθ/α(σ0=F/(hb)), ε(x) is the tensile strain in the viscoelastic tape (assumed to be homogenous along the thickness because the tape is thin), and the tensile strain can be evaluated from the stress σ(x) in the tape through its constitutive model as follows:
(27)
Fig. 4
An illustration of the peeling process in the presence of relative sliding at the interface. τ and σ are the shear stress at the interface and the tensile stress in the tape, respectively.
Fig. 4
An illustration of the peeling process in the presence of relative sliding at the interface. τ and σ are the shear stress at the interface and the tensile stress in the tape, respectively.
Close modal
Here, the implicit assumption is that the tensile deformation is small, allowing the velocity in the peeling arm to be approximated as a constant. J(t) is creep compliance, and the stress in the tape, as shown by the black curve in Fig. 4, is expressed as follows:
(28)

By substituting Eqs. (28) and (27) into Eq. (26), the dissipation from slip can be evaluated.

Consider the architecture in which all components are elastic. The strain in the tape, as shown by the red curve in Fig. 4, can be calculated as follows:
(29)
Then interfacial friction dissipation from Eq. (26) can be evaluated as follows:
(30)
Equation (30) reveals that the interfacial friction dissipation is a function of peeling angle θ. The energy balance equation Eq. (2) can be simplified as follows:
(31)
At steady state, the change in the strain energy d(Usta+Ustp) for a given peel distance da can be evaluated as follows:
(32)
Combining Eqs. (31) and (32) gives the following equation:
(33)

Clearly, Eq. (33) is unable to be applied to the case when the peeling angle is close to zero. The limitation can be attributed to the oversimplified assumption that interfacial shear stress τ is uniformly distributed in a local area α<x<0.

The normalized applied force F/(Γb) is illustrated in Fig. 5. Here, three cases are considered, each represented by different curve colors: soft tape (Eh/Γ=104) with red curves, semi-rigid tape (Eh/Γ=1) with black curves, and rigid tape (Eh/Γ=104) with blue curves. The linear elastic adhesive system is analyzed under two conditions: with interfacial shear stress, represented by solid curves using Eq. (33), and without interfacial shear stress, represented by dash curves using Eq. (4) from Kendall [27].

Fig. 5
Comparative analysis of normalized external force in an adhesive structure with (solid curves) and without (dashed curves) interfacial friction dissipation. Three tape stiffnesses are illustrated: soft (Eh/Γ=10−4), semi-rigid (Eh/Γ=1), and rigid (Eh/Γ=104).
Fig. 5
Comparative analysis of normalized external force in an adhesive structure with (solid curves) and without (dashed curves) interfacial friction dissipation. Three tape stiffnesses are illustrated: soft (Eh/Γ=10−4), semi-rigid (Eh/Γ=1), and rigid (Eh/Γ=104).
Close modal

Interestingly, our analysis reveals that Eqs. (4) and (33) yield nearly identical results for a rigid tape Eh/Γ=104, evidenced by the overlapping blue solid and dashed curves (Fig. 5). However, as the tape becomes more flexible, the disparity between the outcomes of Eqs. (4) and (33) becomes pronounced—particularly evident in the case of extremely soft tape (red curves, Fig. 5). In such scenarios, the equation that omits interfacial dissipation predicts a relatively stable, small normalized peel force (red dashed curve). In contrast, equations that account for interfacial shear stress offer distinctly different forecasts, demonstrating a significant dependence of normalized peel force on the peel angle.

3.4 Discussion of Fracture Energy.

In this subsection, the fracture energy (Γ in Eq. (2)) is discussed in detail. The fracture energy Γ is the energy required to create a unit area of the new interface, expressed as follows:
(34)
where Γ0 is the interfacial intrinsic fracture energy, which represents the energy required to create a unit of area of a new interface without energy dissipation; ΓD is the energy dissipated due to the failure of the interface through processes such as viscoelasticity, plasticity, and viscoplasticity; and Γ0 is rate independent and depends on the type of interface interaction. Notably, Γ0 is often higher than the thermodynamic work of adhesion Wa, especially in the context of peeling structures that incorporate polymers. This discrepancy is largely attributed to the critical contributions of chain breakage or chain pull out, which are key mechanisms in such systems [4144]. Generally, the interfacial intrinsic fracture energy is given as follows:
(35)
where Σ0 is the areal density of the bridging polymer, n is the number of bonds per chain, and Ux is the energy stored in a bond before breaking when bridging chains that span the interface are chemically bonded at both ends [4244], or the energy to pull out the chain when the bridging chains that cross the interface are grafted on one material and remain free at the other end [41]. When the two separate surfaces are connected by van der Waals interactions, the interfacial intrinsic fracture energy can be referred to as the thermodynamic work of adhesion [45].
The thermodynamic work of adhesion Wa of two materials (material A and B) can be evaluated using Dupre's equation [46,47]:
(36)
where σs is the surface energy of material A, σsl is the interfacial tension between the two materials, and σl is the surface energy of material B. While this equation offers a broad perspective, it may overlook the complexity of specific molecular interactions. To address this, the extended formulation was developed by Owens and Wendt [46] as follows:
(37)
where γ1D and γ2D are the dispersive components of the surface energies of the two materials, respectively, and γ1P and γ2P are the polar components of the surface energies of the two materials, respectively.
The energy dissipation ΓD from the interface during crack growth is responsible for the rate dependence of the fracture energy. Many researchers have demonstrated that the interfacial fracture energy is proportional to the interfacial intrinsic fracture energy [4750], which can be expressed as follows:
(38)
where ϕ is a dimensionless function representing the dissipative energy and T is the temperature. Equation (38) is straightforward to understand as higher intrinsic fracture energy results in more significant deformation of the adhesive structure before failure, leading to increased energy dissipation. Therefore, higher intrinsic fracture energy usually corresponds to higher dissipation [51,52]. In some cases, ϕ can be expressed in a simple power law of form [49,53,54]:
(39)
where k is the parameter related to the temperature and thickness of the viscoelastic tape.
Many different mechanisms are applied to interpret ΓD, such as thermally activated bond dissociation, activated rate theory of rubber friction, and slippage of polymer melts on low energy surfaces [41]. For thermally activated bond dissociation mechanisms, the fracture energy can be expressed as follows [41,5557]:
(40)
where α and β are the coefficients related to the thermally activated energy and the size of the interface that dissipates energy. Vp is the velocity of propagation of the peel front. In addition, if the failure occurs cohesively in the adhesive, the fracture energy is a material parameter of the adhesive. Here, the fracture energy can be directly measured by uniaxial tensile tests.

3.5 Energy Terms Associated With the Soft Substrate.

For many peeling systems, the steady-state strain energy in the substrate is negligible [34,55,58,59]. Here, the primary concern is the dissipation energy dUds. The theoretical approaches reported in the literature can be classified as either thick or thin substrate cases.

3.5.1 The Thick Substrate.

Analytical determination of the deformation field in thick substrates is quite complicated. Only a few researchers have attempted to tackle this problem, and in most cases, oversimplified assumptions are necessary to obtain analytical expressions. For instance, Pierro et al. [34] and Afferrante and Carbone [58] analytically evaluated the dissipation of the viscoelastic substrate by assuming that the stress near the peel front is uniformly distributed in a local region of the attached area. They assume that uniform traction is applied over a segment equal to the tape thickness. This assumption is oversimplified and its validity is debatable, as discussed by Ciavarella et al. [60], because the actual interaction zone between the tape and the substrate could be either smaller or larger than the tape thickness depending on many other factors such as friction shear stress. Substrates with a single relaxation time [58] and multiple relaxation times [34] have been considered. Here, the time-dependent displacement on the substrate surface is calculated, and dissipation from the substrate can be evaluated. Subsequently, the authors obtain relationships between the work of adhesion, peeling force, peeling velocity, and Young's modulus of the peeling arm and the substrate. This relation can be reduced to the classical Kendall equation [27] under extremely fast (V) and slow (V0) peeling rates and can be further reduced to the classic Rivlin solution [3] when the substrate is rigid.

Menga et al. [59] consider the V-shaped periodic double peeling based on the energy-based analysis. Here, the authors consider the peeling of an elastic film from an elastic substrate. For the V-shaped doubled peeling structure, the elastic energy stored in the substrate cannot be neglected. Therefore, both the strain energy of the peeling arm and the substrate are considered. To evaluate the strain energy of the substrate, they adopt the analytical solution of the displacement of the soft substrate under uniformly distributed tractions [61]. Perrin et al. [62] obtained the relation between the peeling force and the peeling velocity. They analytically calculated the dissipation of the soft substrate within the limit of the small deformation. In particular, they obtained the relations between dissipation energy and the loss modulus of the soft substrate.

Zhu et al. [55] evaluated the viscoelastic dissipation energy from the soft substrate by using the same approach as Afferrante and Carbone [58]. The authors extended Afferrante's theory to consider the rate dependence of the work of adhesion. In particular, the authors regard the interfacial debonding as the thermally activated rupture process of the polymer chains connecting the two surfaces [41,56,63], and they analytically calculated the corresponding work of adhesion.

3.5.2 The Thin Substrate.

The stress and deformation in thin substrates have been extensively studied [25,26,6471]. Theoretically, by determining the stress and deformation fields, the energy terms associated with the substrate can be calculated, enabling the energy-based analysis. However, in practice, many researchers analyze the failure of the adhesive with thin, soft substrates from a stress-based perspective. This approach is taken because the peel front fracture criteria can be simplified: failure is assumed to occur when the stress or deformation at the peel front exceeds a critical value. In the next section, we will delve into stress-based analysis in greater detail.

4 Stress-Based Analysis

In this section, we will focus on adhesives with a thin, soft substrate. Several researchers independently developed the theory to describe the peeling mechanics of the elastic tape peeling from a thin elastic substrate [6670,72]. The basic idea of these studies is similar; the tape is considered a bending beam, the stress and deformation field in the soft substrate of the bonded area is solved, and the failure criterion is set as the stress/strain reaches a critical value. Specifically, Kaelble [66] assumed that the soft substrate undergoes in-plane shear and uniaxial deformation. Dillard [67] used a similar idea, and the author considered the bending of the plates on incompressible elastic foundations. The elastic foundations are assumed to be under a uniaxial deformation state. More recently, this theory was applied to study the peeling process of finite-length tape on incompressible elastomeric foundations [69]. Plaut et al. [68] investigated the peeling of finite-length beams on elastic foundations. The author modeled the substrate as independent springs. The beam was modeled as inextensible, and they only considered the bending of the beam. In addition, a series of works [26,70,71] theoretically calculated the deformation and stress field in thin, soft substrates by solving the 2D equilibrium equation. The detachment criterion is defined as the condition when the stress in the substrate reaches a critical value.

Two categories can be identified based on the deformation of the soft substrate: those with small (see Fig. 6(a)) and large deformations (see Fig. 6(b)). In practical applications, in the first case, the tape consists of both backing and adhesive materials. The soft substrate typically corresponds to a soft release liner, where interfacial adhesive failure usually occurs due to the small interfacial fracture energy; here, the release liner undergoes little deformation. In contrast, in the second case, the soft substrate can undergo large deformation due to the stronger interfacial fracture energy and/or low modulus of the adhesive. Here, different types of failure can occur depending on loading conditions, material, and interface properties. For example, peeling a bandage from the skin typically represents adhesive failure, where the tape shown in Fig. 6(b) is the bandage, i.e., backing and adhesive, and the substrate is the skin. Conversely, many of the cases involving highly deformable substrates detailed later in this article involve substrate failure, where failure occurs within the highly deformable substrate (light blue) as shown in Fig. 6(b).

Fig. 6
An illustration of peeling from two types of soft substrates: (a) thin, soft release liner, where the tape consists of both backing and adhesive materials, and (b) highly deformable substrate. Here, the tape consists of both backing and adhesive materials for adhesive failure, which occurs between the tape and the substrate. Other instances of failure may instead occur within the deformable substrate.
Fig. 6
An illustration of peeling from two types of soft substrates: (a) thin, soft release liner, where the tape consists of both backing and adhesive materials, and (b) highly deformable substrate. Here, the tape consists of both backing and adhesive materials for adhesive failure, which occurs between the tape and the substrate. Other instances of failure may instead occur within the deformable substrate.
Close modal

The stress-based analysis is fundamentally grounded in a similar procedure, specifically the stress balance method. Therefore, only one seminal work, serving as a representative example for each of the two types of structures, is discussed herein. In particular, an overview of cleavage stress analysis [25,66,73,74] is provided for the first structure (Fig. 6(a)), while the second structure (Fig. 6(a)) is discussed in terms of the balance of momentum of the tape, as described in theoretical approaches by Derail et al. [75,76] and Li et al. [77].

4.1 Peeling From a Soft Substrate Under Small Deformation.

Cleavage stress analysis is specifically developed for peeling a flexible backing from a deformable linear elastic substrate with small deformation [25,66], as shown in Fig. 7. Given its foundation in the principles of deformable elastic behavior, this theory finds natural applicability across a broad range of adhesives with soft elastic substrates. In this subsection, the terms “soft substrate” and “soft release liner” are used interchangeably, unless specified. The peeling front is denoted as point A. m denotes the bending distance between point A and the peeling force. The thicknesses of the tape and soft release liner are denoted as h and hl, respectively. An element dx in the bonded area is considered (see the inset in Fig. 7). By applying an equilibrium condition for the external moments and forces for the element, the following equations are derived, following Ref. [66]:
(41)
(42)
(43)
where T is the tensile force acting on the tape, and q and f are the shear and tensile force applied by the soft release liner, respectively. M is the total bending moment, and s0 is the shear force applied on the tape.
Fig. 7
An illustration of tape peeling from a soft release liner. m denotes the bending distance between the peel front and the peeling force. An element dx in the bonded area is magnified for illustration of the equilibrium of the external moment and forces. T and s0 are the tensile force and shear force in the tape, respectively; M is the total bending moment applied to the tape; q and f are the shear forece and tensile force applied by the soft release liner, respectively; and u is the shear displacement of the soft release liner.
Fig. 7
An illustration of tape peeling from a soft release liner. m denotes the bending distance between the peel front and the peeling force. An element dx in the bonded area is magnified for illustration of the equilibrium of the external moment and forces. T and s0 are the tensile force and shear force in the tape, respectively; M is the total bending moment applied to the tape; q and f are the shear forece and tensile force applied by the soft release liner, respectively; and u is the shear displacement of the soft release liner.
Close modal

4.1.1 Shear Stress.

First, the shear force q is evaluated. The shear stress on the soft release liner can be expressed as follows:
(44)
where Gl is the shear modulus of the soft release liner, and u is the shear displacement of the soft release liner. Substituting Eq. (44) into Eq. (41) gives the following equation:
(45)
Note that the tensile displacement in the tape is the same as the shear displacement of the soft release liner at x, attributable to the intimate contact between the tape and the liner. Therefore, for the linear elastic tape with Young's modulus E, the tensile displacement satisfies:
(46)
Combining Eqs. (45) and (46) gives
(47)
Solving Eq. (47), the expression for the tensile force is given as follows:
(48)
where
(49)
Since T()=0 and T(0)=Fcos(θ), then tensile force can be determined as follows:
(50)
The shear stress between the soft release liner and the adhesive is given by S=q/(bdx). Combining Eqs. (41) and (50) gives:
(51)
Equation (51) can be written as follows:
(52)
where SA=αFcos(θ)/b is the shear stress at peel front A. The total shear force Fs between the tape and the soft release liner is given as follows:
(53)
The bending moment at the peel front caused by shear stress is given as follows:
(54)

4.1.2 Cleavage Stress.

Next, we evaluate the tensile cleavage stress (σc=f/(bdx)) between the tape and the soft release liner. The derivation of the cleavage stress necessitates the use of bending equations, as summarized in Ref. [66]. For an elastic tape, the bending equation is as follows:
(55)
The bending deformation in the bonded area of the tape and the soft release liner is small, approximately dϕ/ds=d2y/dx2(|dy/dx|1). Then Eq. (55) can be rearranged as the Euler–Bernoulli beam theory:
(56)
Substituting Eq. (43) into Eq. (56) gives the following:
(57)
From Eq. (42), the following equation is obtained:
(58)
Combining Eqs. (57) and (58) gives the following:
(59)
with the following boundary conditions:
(60)
where
(61)
and MA=Fm is the bending moment at the peel front (see point A in Fig. 7). The distance m between point A and the peeling direction is defined in Appendix  A, Eq. (A3). Equation (59) is derived based on the assumption that bending of the tape is mainly caused by an externally applied force.
By solving Eq. (59), the displacement of the tape in the bonded area is given as follows:
(62)
The coefficients A and B are expressed as follows:
(63)
(64)
The tensile cleavage stress between the tape and the soft release liner is expressed as follows:
(65)
where σA is the stress of the peel front, evaluated as follows:
(66)
(67)
where m is the bending distance between point A and the peeling force, which is given in Appendix  A. Gent and Hamed [78] indicated that K1 satisfies the small bending deformation assumption.
Fig. 8
Variation of (a) dimensionless cleavage stress and (b) shear stress with dimensionless distance from the peel front. Higher α and β values correspond to the increased liner rigidity, leading to a more concentrated distribution of cleavage and shear stresses near the peel front. This concentration results in stresses dissipating over shorter distances, underscoring the critical role of liner stiffness in stress behavior.
Fig. 8
Variation of (a) dimensionless cleavage stress and (b) shear stress with dimensionless distance from the peel front. Higher α and β values correspond to the increased liner rigidity, leading to a more concentrated distribution of cleavage and shear stresses near the peel front. This concentration results in stresses dissipating over shorter distances, underscoring the critical role of liner stiffness in stress behavior.
Close modal
The total tensile force Ft between the tape and the soft release liner can be evaluated as follows:
(68)
Based on the force equilibrium along the y direction, the following relationship is established as follows:
(69)
The bending moment caused by the tensile stress can be calculated as follows:
(70)

Figure 8 illustrates the relationship between normalized shear stress (Eq. (52)) and cleavage stress (Eq. (65)) at the interface of the soft release liner and adhesive, and their distance from the peel front. It is observed that both cleavage stress and shear stress diminish to zero as the distance from the peel front increases. Specifically, cleavage and shear stresses disappear at distances corresponding to βx and αx values of approximately −4 and −6, respectively. A notable distinction is observed in the behavior of these stresses: the cleavage stress exhibits oscillatory behavior before dissipating, whereas the shear stress decreases monotonically as distance increases. It is important to note that the parameters α and β are indicative of the stiffness of the release liner, with higher values suggesting increased rigidity. Consequently, enhancing the rigidity of the liner results in a more localized concentration of both cleavage and shear stresses, which then dissipate at shorter distances from the peel front.

Fig. 9
An illustration of peeling from a soft, highly deformable substrate. An element ds in the peeling arm is magnified.
Fig. 9
An illustration of peeling from a soft, highly deformable substrate. An element ds in the peeling arm is magnified.
Close modal

4.1.3 Combined Stress Mode.

In the previous two subsections, we discussed the cleavage tensile stress and the shear stress independently. Now, by using the force equilibrium equation, we can derive the relationship between the steady-state peel force and the combined stress (i.e., a combination of cleavage tensile stress and shear stress). Considering the force equilibrium, the following equation holds [66]:
(71)
The total moment at the peel front caused by the bond is MA=MAt+MAs. Combining Eq. (A4) gives:
(72)
Reorganizing this equation gives
(73)

Here, the cross section is considered rectangular, that is, I=bh3/12.

4.1.4 Comparison of Energy-Based and Cleavage Stress Analyses.

Consider an elastic adhesive where the kinetic energy and the extensional deformation of the tape are trivial, Eq. (3) holds and is rewritten here as follows:
(74)
Based on cleavage stress analysis (Eq. (73)) (tensile stress dominated), the following equation holds:
(75)

Equations (74) and (75) give the same dependence of peel force on the peeling angle. Note that it may seem that Eq. (75) predicts that the peel force will depend on the release liner thickness. However, Gent and Hamed [79] pointed out that σA is the average stress along the thickness of the soft release liner at the peel front. Therefore, σA is not a constant value, and it may be inversely proportional hl1/2. As a result, both equations are independent of the liner thickness. Interestingly, comparing Eqs. (74) and (75) gives the equation ΓhlσA2/2El, which represents the elastic energy stored in the thickness of the soft release liner near the peel front.

4.1.5 Remarks on the Cleavage Stress.

The cleavage stress method was initially developed for adhesives that feature an elastic tape and release liner, but this method can be extended to characterize structures with nonelastic properties. For instance, to capture the peeling behavior of structures with a viscoelastic release liner or tape, the elastic constitutive equations such as Eqs. (44), (46), and (55) could be replaced by the corresponding constitutive equations such as viscoelastic constitutive equations. The only premise for this approach is that the deflection deformation in the bonded area of the tape and the soft release liner should remain small (see Eq. (56)). This method can also be extended to consider the finite deflection deformation in the bonded area of the tape and the soft release liner by using the general bending equation, Eq. (6). In this case, the soft substrate experiences large deformation during the peeling process, as discussed in Sec. 4.2.

4.1.6 Advances in the Theory of Soft Substrates Under Small Deformation.

Spies [64] and Bikerman [65] established the basis for the cleavage stress method through their investigation of the detachment mechanics of an adhesive consisting of a flexible peeling arm adhered to a rigid plate with linear elastic adhesive. Based on this, Kaelble [25,66] developed the cleavage stress approach, and this theory was subsequently confirmed by experiments [74]. The basics of the cleavage stress analysis are provided in Sec. 4.1.2. This theory combined with the rate–temperature superposition principle is used to describe the peel force of the adhesive structure with a viscoelastic adhesive interlayer [73,80]. Chen and Flavin [81] extended the cleavage stress by considering the elastic–plastic tape and the elastic–plastic soft substrate. The authors found that the fracture energy is the key parameter instead of the specific constitutive relations.

Moidu and Sinclair [4,5] considered the peeling process based on both the energy and cleavage analyses. Here, the deformation in the detached area is analyzed based on the general bending theory [30,82]. The deformation in the attached area is calculated based on the cleavage stress analysis. The authors considered the elastic–plastic tape peeling from the elastic soft substrate. Energy analysis was employed, and the corresponding strain energy and the dissipation energy from plasticity were calculated from the calculated deformation. Ghatak [26] investigated the elastic flexible plate peeling from the elastic substrate with varied modulus along the x direction. The author applied the stress equilibrium equation (balance of linear momentum) to the material element in the soft substrate and the bending equation to the flexible tape. It was assumed that the plate deflection was small in the attached area of the plate and the soft substrate. The peel front was then determined by the failure criterion that the adhesive breaks at a constant critical stress. Finally, Villey et al. [83] evaluated the fracture energy (Γ) using a stress-based analysis. The peeling force, derived from the cleavage stress (Eq. (69)), is employed to assess the fracture energy. The authors analytically separated the angular dependence of the fracture energy from the dependence on peeling velocity, finding that the angular dependence can be explained through the stress-based analysis. Additionally, they found that accounting for large deformations is necessary to explain the high fracture energy values observed in PSAs.

4.2 Peeling From a Soft Substrate Under Large Deformation.

The structure of the soft substrate under large deformation is shown in Fig. 6(b). As aforementioned, various types of failure can occur. This highly deformable substrate experiences the maximum stretching at the peel front and gradually decreases toward the left of the peel front. The point where the stretch is zero, denoted by O, serves as the origin of the Cartesian coordinate system (Fig. 9). Note that the substrate on the left of point O experiences compression, and the influence from the left side of O on point O can be represented by a force F0 and a moment Mf. The soft highly deformable substrate is assumed to be under uniaxial deformation, which can be illustrated in a series of springs (see Fig. 9).

4.2.1 Momentum Balance on the Tape.

Applying a linear momentum balance to the tape on the right-hand side of O gives the following [75]:
(76)

To obtain the peel force F, the value F0, the constitutive model of the deformable substrate σa, the length of the peeling zone l, and the shape of the bent arm y(x) need to be evaluated. Assuming that the bent arm is circular, the theoretical method can be simplified (see Appendix  B). A more rigorous method is introduced to determine the shape of bent arm y(x).

Considering the linear elastic peeling arm, the bending equation of the material element (see the inset of Fig. 9) at the position with arc length s is given as follows:
(77)
This equation can be written in the Cartesian coordinate system as the large deflection theory, summarized in Ref. [84]:
(78)
The peeling zone is depicted in Fig. 9. The bending moment at position x is given as follows:
(79)
The moment Mf is assumed to be vanishing [84]. At x=0, the following boundary conditions hold
(80)
At x=l,
(81)

Here are two unknown parameters F0 and l, which can be numerically determined by using Eqs. (79)(81). Note that σa is determined by the constitutive model of the deformable substrate material, while δmax depends on the failure conditions, and an example of the failure condition is discussed in Appendix  B in Determination of the Angular Position of the Failure Locus θM section.

4.2.2 Remarks on the Angular Momentum-Based Approach.

While initially developed for elastic tape (Eq. (77)), this approach can be extended to other material systems. Elastic–plastic [20] and viscoelastic tapes [33] can also be analyzed using the corresponding constitutive models. Furthermore, if the traction–separation law is employed for the highly deformable substrate, this theoretical method is consistent with the cohesive zone model [84].

4.2.3 Advancements in the Theory of Substrates Undergoing Large Deformation.

Cleavage stress analysis is valid when deformation in the bonded arm is small. Niesiolowski and Aubrey [85] extended the cleavage stress to consider the influence of filamentation by analyzing the peeling profile in the peeling zone with large deformation. Derail et al. [75,76] considered peeling mechanics that can be applied to a highly deformable, viscoelastic substrate by assuming that the peeling profile is circular and the highly deformable substrate in the peeling zone undergoes uniaxial deformation. The authors considered both cohesive and interfacial adhesive fractures by using the corresponding failure criteria (Appendix  B).

Karwoski [86] explored the mechanics of tape peeling from the skin, summarizing several models where the force and moment equilibrium equations were utilized to establish the governing equations. To streamline the analysis, the soft deformable substrate, representing skin tissue, was modeled as a single spring—a simplification from a more complex multispring system. Additionally, both the skin surface and the tape were modeled as flexible beams. Subsequently, Plaut [87] delved into similar investigations of tape peeling behavior from the skin. This model partitioned the structure into two distinct segments: segment 1, where the tape is adhered to the skin, and segment 2, where the tape is detached. In segment 1, the mechanical behavior is primarily governed by the tape backing, which is modeled as a flexible beam. Conversely, in segment 2, the skin surface is represented as a truss (exhibiting no bending), and the soft substrate is modeled as a series of springs.

Zhang and Wang [88] examined bending in the attached and detached areas of the tape. The authors considered cohesive failure within the highly deformable substrate. Here, the Maxwell solid constitutive relation is applied to the substrate, and the detached location l between the substrate and the peeling arm is determined by the failure criterion that the substrate fibril breaks at a constant critical stretch ratio.

Poulard et al. [89] theoretically investigated the deformation of a substrate with pillar patterns. The authors assumed that all strain energy in the pillars would dissipate during the peeling process, which will increase the peeling force. Here, failure criteria similar to Johnson–Kendall–Roberts and Derjaguin–Muller–Toporov theory are selected, where the tape will detach from the substrate as the force in the pillar reaches a critical value.

Li et al. [77] derived the peel force based on a balance of linear momentum. The authors use a series of springs to represent the highly deformable substrate layer and tape. Force equilibrium was utilized, and bending of the tape was ignored. They found that the critical strain/stress of the substrate when the adhesive failure occurs is not constant and depends on the peeling angle.

Finally, He et al. [84] used a soft hydrogel as the highly deformable substrate and applied the theoretical framework in Sec. 4.2.1 to capture the debonding resistance of the hydrogel, with the further assumption that Mf=0, i.e., y(0)=0. In their research, cohesive failure always occurs. Thus, the authors directly measured the constitutive relation of the deformable substrate σa with corresponding δmax. In particular, they determined the traction–separation law of the deformable substrate by using the butt-joint test [90].

4.3 Summary of Theoretical Models.

The theoretical model section of this review critically assesses the theoretical underpinnings of peeling mechanics, particularly within soft adhesives that interface with soft substrates, emphasizing both energy and stress analyses. Distinctive failure criteria characterize each method. In the energy-based analysis, the peel front propagates once the external energy input suffices to generate a new surface. Conversely, the stress-based analysis dictates that the peel front moves forward when the stress or deformation at the peel front reaches a critical threshold.

For the energy analysis method, different energy terms are discussed. This analysis encompasses energy terms in both the tape and substrate, kinetic energy, energy dissipated through interfacial slippage, and fracture energy at the interface. The energy terms related to the tape—often modeled as a flexible beam—are comparatively straightforward to assess. This review covers the energy terms arising from the bending and tensile deformation of the peeling arm, particularly highlighting the elastic, elastic–plastic, and linear viscoelastic tape models. The roles of the kinetic energy and the energy dissipated from interfacial slippage are discussed, with seminal works introduced in Sec. 3.3 serving as prime examples. Moreover, the intrinsic fracture and the associated energy dissipation at the interface, along with the rate and temperature dependence of the fracture energy, are reviewed. Though the energy terms connected to the substrate present challenges, these energy terms can be evaluated with certain assumptions. For instance, for soft, thick substrates, oversimplifications are utilized, and for soft, thin substrates, the analysis stays constrained to small deformations. In the case of soft, thin substrates, energy terms can be derived from their stress and deformation fields. Practically, stress analysis typically dominates the analysis for structures incorporating thin substrates.

On the flip side, stress analysis—which predominantly focuses on the stress and deformation dynamics within a structure—is mainly pertinent to structures with thin, soft substrates. This specificity arises because the assumption of a thin substrate allows for a feasible mathematical derivation; without this assumption, the mathematical approach becomes intractable. This review covers the nuances of soft substrates experiencing both minor and substantial deformations, typically corresponding to release liners and adhesives. Seminal works by Kaelble [25,66] on soft substrates with minor deformations are dissected in depth. The application of the linear Euler–Bernoulli beam theory to the bonded area of the tape is explored, resulting in the analytical derivation of cleavage and shear stress in the elastic substrate. Notably, both cleavage stress and energy analyses reveal congruent angle dependencies. As for soft substrates with substantial deformations, large deflection theory becomes indispensable, particularly for capturing tape deformation. To solve the governing equations of peeling mechanics, one must first extract and calibrate the constitutive model of the soft adhesive and the corresponding failure conditions from experiments.

This review also highlights recent advancements in the energy and stress analysis of structures with both thin and thick substrates, as summarized in Table 1. In Table 1, a summary of various models is presented, where T and S in the rows and columns represent the tape and substrate, respectively. The acronyms used within the table are as follows: LE for linear elastic, NE for nonlinear elastic, EP for elastic–plastic, and LVE for linear viscoelastic. The terms WB and SB indicate weak bonds and strong bonds, respectively, denoting the substrate's response to small deformations under weak bonds and large deformations under strong bonds. Additionally, TnS and TkS are used to differentiate between thin and thick substrates.

Table 1

Summary of references for the soft peeling system with the thin and thick substrate

Peeling structureS-LES-NES-LVE
WBSBSBWBSB
T-LETkS[59][34,55,58,60,62]
TnS[25,26,6471][85] [77] [87] [86][84][73,83][88]
T-EPTnS[4,5,81][75,76]
T-LVETnS[89]
Peeling structureS-LES-NES-LVE
WBSBSBWBSB
T-LETkS[59][34,55,58,60,62]
TnS[25,26,6471][85] [77] [87] [86][84][73,83][88]
T-EPTnS[4,5,81][75,76]
T-LVETnS[89]

Note: T, soft tape; S, soft substrate; LE, linear elastic; NE, nonlinear elastic; EP, elastic–plastic; LVE, linear viscoelastic; WB, weak bonds; SB, strong bonds; TnS, thin substrate; TkS, thick substrate.

It is important to note that many theoretical works have been deliberately excluded from this review, as the focus is tightly bound to peeling systems with soft substrates. Consequently, adhesive structures of different types are not covered in this review. For instance, certain studies explore scenarios such as tape peeling from thin beams or trusses. Notably, Shen et al. [91] consider peeling a thin film from a soft cantilever substrate by the energy balance analysis. In this work, bending of the cantilever substrate is analytically solved to determine the peeling angle and then the model of Kendall [27] is employed to calculate the peeling force. Peeling of PSAs from the human skin, in which the human skin is modeled as a reversible truss (elastic foundation and bending have not been incorporated), has been investigated by several authors [9294].

In addition, theoretical works pertinent to rigid substrates have also not been incorporated into Table 1. For the theoretical works related to rigid substrates, seminal contributions were made by Rivlin [3] and Kendall [27]. A selection of additional works, though not exhaustive, will be briefly mentioned here. Loukis and Aravas [95] explored linear viscoelastic tape peeling from a rigid substrate, while Christensen and McKinley [96] studied linear viscoelastic and nonlinear viscoelastic tape peeling from a rigid substrate. Similarly, exploration by Afferrante et al. [97] into the peeling of elastic thin films from a rigid substrate, investigation by Chen et al. [33] into the flat and patterned viscoelastic tape peeling from a rigid substrate (as detailed in Sec. 3.1), and the study by Peng et al. [54] on the viscoelastic thin tape peeling from the rigid substrate (discussed in Sec. 3.4) have been consciously omitted. Additional research studies such as that by Zhu et al. [98], who investigated visco-hyperelastic tape peeling from the rigid substrate at a zero-degree peeling angle, and Ceglie and Menga [8,99], who explored the peeling behavior of thin viscoelastic tape from a rigid substrate, have also been excluded. Similarly, works by He et al. [32,100], examining hyperelastic tape peeling from a flat rigid substrate, are not covered in this review.

5 Specific Applications of Peel Mechanics of Deformable Tape and Substrate

Peeling of a soft adhesive from a soft substrate is a common phenomenon encountered in various applications. For example, when removing an adhesive label from an elastomer-coated release liner before applying it to another substrate, achieving effortless removal without compromising the adhesive strength becomes crucial. Similarly, for skin adhesives or transdermal patches, the process of painless removal from human skin is of utmost importance. Additionally, when assessing the interfacial strength of multilayer packaging films, the peeling process is a valuable tool. Consequently, it becomes essential to develop a comprehensive understanding of the underlying mechanics of this peeling process and identify the diverse fields in which this knowledge finds significant applications. These endeavors play a pivotal role in the successful material development, precise characterization techniques, and accurate interpretation of experimental data. In this section, we aim to establish a connection between the previously discussed theories of peel mechanics and practical applications commonly experienced by industrial researchers and customers. We will demonstrate how the theoretical foundations and testing methodologies can effectively guide the creation and understanding of soft materials, both within and beyond the confines of the laboratory.

5.1 Peeling Soft Adhesives From Soft Release Liners.

Soft adhesive tapes or pressure-sensitive adhesives are extensively employed for surface bonding in various aspects of daily life, playing indispensable roles in a broad spectrum of industries, including electronics, construction, automotive, and healthcare [101,102]. However, these adhesives are often designed to exhibit high tackiness upon contact to enable rapid bonding, making them prone to accumulating contaminants on the adhesive surface. To safeguard the adhesive surface from contamination prior to application on a substrate, release liners can be utilized. These liners consist of plastic or paper sheets, which can be coated with release coatings if necessary to meet the required release performance. The adhesive and liner layer can be two separate entities, most commonly for labels. For many self-wound tapes (e.g., packing tape), the adhesive and liner layers are not separate, with one side coated with an adhesive and the other side either coated or uncoated with a release coating. In this context, two crucial performance parameters are used to evaluate the effectiveness of release coatings in relation to the PSA: release force and subsequent adhesion strength (SAS) [103106]. The release force measures the ease with which the PSA can be peeled off from the release coating or backing or vice versa, while the SAS represents the adhesive strength of the PSA upon removal from the release coating or backing and its subsequent attachment to a target substrate. The release performance parameters can be easily measured by mechanical testing equipment. Although the numerical aspects of release performance and visual observations by an operator can offer prompt information on failure modes (e.g., cohesive, or adhesive failure) and crack dynamics (e.g., stick-slip), they often fall short in providing insights into the fundamental origins of the release performance. Understanding the fundamental factors that influence global release performance is essential for developing new adhesives or release coatings, as well as optimizing and troubleshooting existing ones.

5.1.1 Release Mechanisms and Technical Approaches.

One of the most employed mechanisms and approaches in the development of release coatings involves reducing the surface energy to enhance repellency against the PSA and enable facile release. For elastic tape, the release force can be determined by the interfacial fracture energy via Eq. (4). More generally, the correlation between release force and fracture energy can be determined through the energy-based analysis (Sec. 3). On obtaining the relationship between release force and fracture energy, the relationship between surface energy and release force can be derived via Eq. (38). Using elastic tape as an example, the release force increases with an increase in surface energy, and vice versa (combining Eqs. (4) and (38)). In practical terms, the minimum release force, corresponding to the smallest fracture energy, can be achieved by relying solely on van der Waals interactions, specifically dispersion forces. Consequently, stronger interactions such as hydrogen bonds, acid–base interactions, covalent bonds, or the interdigitation of short chains, along with interfacial entanglements of sufficiently long chains, result in the amplification of viscoelastic energy dissipation and higher release forces.

In addition, the stress-based analysis presented in Sec. 4.1 offers a framework that could be potentially adapted for analyzing the peel mechanics of the PSA-release liner system. However, this section models the tape as a homogeneous elastic entity, a simplification that diverges from more complex real-world scenarios. In practice, tape comprises two distinct materials: a backing material, which may be modeled as an elastic material, and an adhesive material, characterized by pronounced viscoelastic properties. Consequently, caution must be exercised when applying this model. Specifically, when treating the inhomogeneous tape as a singular elastic material with an equivalent modulus, the model applicability becomes restricted to peel velocities within a certain range where the adhesive loss modulus remains negligible.

Interestingly, the relationship between surface energy and release force as implied in Eq. (38) was found to break down when peeling an organic PSA from a polydimethylsiloxane (PDMS) coating. To investigate this phenomenon, peel experiments were conducted using PDMS, hydrocarbon, and fluorocarbon-grafted coatings [107]. Despite PDMS having a higher surface energy (γ = 22 mN/m) compared to fluorocarbon (γ = 10 mN/m), the release force of the organic PSA was significantly lower for PDMS than for fluorocarbon, at a given peel velocity. Furthermore, each coating exhibited distinct peel front patterns, with PDMS showing the smoothest pattern and the lowest peel force, while fluorocarbon demonstrated a highly irregular pattern with the highest peel force.

This deviation in release forces relative to surface energy can be attributed to interfacial slippage induced by the high surface flexibility and mobility of PDMS. This slippage was visually characterized by the lateral displacement of fluorescent particles at the interface during peeling [40,107]. The fluorescent particle on PDMS moved approximately 13 µm before encountering the crack, whereas the particle on fluorocarbon only moved 1 µm. These surface dynamics observed in PDMS effectively reduce interfacial shear stress, leading to minimized energy dissipation and low release force. Several researchers have incorporated the dissipation due to interfacial slippage in their models [7,8,40]. In contrast, no interfacial slippage was observed between the organic PSA and fluorocarbon coating, resulting in a high release force due to significant energy dissipation in the adhesive, despite its extremely low surface energy.

The concept of chain flexibility in controlling release forces extends beyond PDMS, as evidenced by investigations on hydrocarbon monolayers (C16) in relation to acrylic PSAs [107]. In the liquid-like state (65% surface coverage), the chains exhibit higher mobility, allowing for more slippage, which results in a reduced release force compared to the crystalline state (100% surface coverage). However, an interesting counterargument regarding the impact of chain flexibility on release force control has also been reported. In a polycarbamate-based organic release coating system with an organic PSA, a high crystalline solid-like surface state through a high density of long alkyl chains (C16) proves advantageous in reducing the release force by suppressing entanglement with the chains of the organic PSA, compared with shorter alkyl chains (C10).

While achieving a lower release force is often a primary goal in industry, avoiding excessively low release forces remains crucial to prevent potential technical issues. For example, during storage or die-cutting into desirable forms, self-delamination of the laminates may occur. Therefore, attaining appropriate release forces based on specific applications and processing requirements is imperative. An effective approach to controlling release force in release coatings such as PDMS involves modifying elastomeric architectures through adjustments in the degree of polymerization and/or branching [108]. This tuning process not only leads to a wide variation in energy dissipation and release performance but also allows for customization of curing conditions, providing a means to achieve the desired release forces, and curing profiles.

Another approach is to add methyl silicates or MQ resins into the elastomer [109,110]. Notably, the nanoscale additives can suppress interfacial slippage in PDMS release coatings, leading to an increase in the release force. The addition of methyl silicates, which serve as high release additives (HRAs), increases the loss modulus in PDMS coatings, reducing chain flexibility and creating a surface freeze effect. This, in turn, results in increased friction, elevated energy dissipation within the sheared adhesive layer, and higher release forces. The impact of HRA on weakening interfacial slippage becomes more pronounced over extended contact times due to an increase in interfacial interactions. In addition to molecular-level chain dynamics, micropatterns on the PDMS surface at a micron scale can significantly influence the release force of the acrylic PSA through elastic deformation of the patterns [89,111].

While release force is a critical parameter for release liners, SAS holds equal importance. By laminating the PSA onto a release liner featuring a surface coated with a thin layer of slippery liquid oils, the release force can be significantly reduced [112]. However, when the PSA is laminated onto a target substrate, adhesion strength may vanish due to the migration of oils onto the adhesive surface. Therefore, careful consideration and evaluation are necessary to balance the reduction of release force while maintaining sufficient SAS.

5.1.2 Characterization of Release Performance.

Release or adhesion performance is significantly influenced both by the materials used and by environmental factors and testing conditions. As a result, evaluating release performance typically entails conducting characterization under various lamination conditions and peeling rates for both the PSA and the release liner, tailored to meet specific application requirements.

One common condition involves adhering a PSA strip onto a prepared release liner with a rubber roller with an approximate weight of 5 kg. Subsequently, the laminate is subjected to extended periods under a heavy weight, either at a room temperature or at an elevated temperature (e.g., 70 °C). Following aging, the laminate is affixed to a release tester, and the PSA or release liner end is pulled at a 180-deg angle at a rate of 0.3 m/min. The removed PSA is then laminated onto a stainless-steel plate and peeled at the same rate to characterize the SAS (ASTM-D6862 [113], D3330 [114], and D6252 [115]). Various ASTM standards for peel tests cater to different application scopes. For instance, ASTM D6252/D6252M is employed to measure the peel adhesion of PSA label stocks. A more thorough review of these various peel tests can be found in the study by Bartlett and Case [19].

Although commonly used, the specified peeling rate above (0.3 m/min) is considered slow for many applications, particularly in cases where peeling of high-volume PSAs from release liners is conducted on an industrial scale. To assess the effectiveness of release liners and meet demanding requirements, a much higher peeling rate, such as 300 m/min, is often selected [105,116118]. This high-speed release can reveal significant deviations in release performance compared to slower peeling rates, depending on the types of PSAs and release liners employed. In particular, at high peeling rates, there is a notable dissipative process that significantly influences the overall release performance. The dissipative behavior primarily occurs within the soft and thick PSA (≥25 µm) rather than the thin release coating (≤1 µm), mainly due to the PSA's larger dissipation factor (tan δ), which is closely linked to the adhesion strength on a target substrate. The dissipation energy from the soft adhesive can be evaluated by using the approach discussed in Sec. 3.1.3, and the specific method can be selected from Table 1.

Interestingly, the tan δ profile of acrylic PSA, obtained through rheological tests at various frequencies, can be qualitatively correlated with the release profile of PDMS coatings at the corresponding peeling rates [108]. Although the contribution of the PSA to release force is apparent, the effects of the degree of polymerization and quantities of MQ resins on the release coatings are also significant. For instance, tightly crosslinked unfilled PDMS coatings or modified PDMS coatings with low quantities of MQ resins exhibit low release forces for organic PSA at high peeling rates [108].

The geometry and mechanical stability of the release coating also significantly impact the release performance and must be extensively characterized, along with the measurement of release force and SAS. The coating thickness is carefully controlled by taking into consideration factors like performance, processing, and cost. To measure the thickness, nondestructive analytical techniques, such as X-ray fluorescence, are often employed. These methods also facilitate rapid characterization of the mechanical strength of cured coatings after they have been exposed to abrasion or solvents, which is critical for practical long-term applications. For example, the release liner is immersed in a specific solvent to characterize the presence of uncured prepolymers that may migrate to the surface of PSA, resulting in reduced SAS. The surface of the release liner is also pressed against a fabric strip under a weight and subjected to multiple rubbing cycles to evaluate the cohesive strength of the release coating.

5.1.3 Market, Prospect, and Opportunity.

Release coatings are micron-thick materials applied and cured onto release liners to provide weak interaction with sticky surfaces and prevent premature and permanent adhesion. Types of release coatings include polyacrylates, carbamates, polyolefins, fluorocarbons, chromium stearate complexes, and silicones. Among them, silicone- or polydialkylsiloxane-based release coatings dominate the market due to their superior performance over other types, such as extra low release force, consistent release performance, chemical and heat resistance, and environmental friendliness [119,120]. The extraordinary performance is mainly due to the unique chemical structure of silicones, resulting in extra low release force that cannot be achieved by other types of release coatings as discussed earlier [40,107,121], making them ideal for use as release coatings. Release liners coated with a release coating are widely utilized in various areas, particularly in the PSA sector where they play a critical role in protecting the sticky PSA layer during manufacture, storage, and facilitating easy delamination and label transfer from liners onto the targeted labeling objects. The typical application of release liners includes labels, tapes, hygiene products, and food/bakery products (Fig. 10). The total global release liner market size is about 56,717 million m2 with steady growth in 2020 [122].

Fig. 10
Typical applications of silicone release coatings
Fig. 10
Typical applications of silicone release coatings
Close modal

One of the crucial development trends in the release coating sector is the enhancement of sustainability in release coating systems. This includes transitioning from solvent-based coatings to emulsion and solventless alternatives, developing release coatings with a reduced carbon footprint, creating recyclable release liners, and more. However, such changes are not without consequences, as alterations in components can impact release performance. To ensure that release coatings can effectively meet these evolving trends and coating performance targets, a profound understanding of the fundamental peeling mechanism is critical. This knowledge is essential for guiding the development direction and ensuring a prompt and successful response to market trends, thereby contributing to the overall improvement of sustainability.

As discussed earlier, the release liner market holds promising prospects and significant opportunities, particularly in the context of research papers that delve into the peeling mechanics of soft adhesives from release liners. This niche area of study plays a pivotal role in enhancing the understanding of adhesion dynamics, which in turn has wide-ranging implications across industries. This review paper focusing on peeling mechanics is poised to contribute valuable insights into the fundamental interactions between soft adhesives and release liners and a fundamental understanding of the delamination mechanism, shedding light on factors influencing adhesion and release performance.

As academia and industries continue to seek efficient and reliable ways to optimize adhesive and release coating performance, this review article provides deeper insights into the peeling mechanism of soft adhesives and opens doors to novel solutions and improved products to meet further needs. For example, the viscoelastic dissipation in the adhesive tape is supposed to play a vital role in release force. Section 3.1.3 discusses the details for the evaluation of the dissipation energy term for linear viscoelastic dissipation energy. In addition, the stress and deformation of a release liner, especially the stress and deformation around the peeling front, can be evaluated by using the method introduced in Sec. 4.1. Note that Sec. 4.1 considers the both tape and release liner as linear elastic materials. However, in practice, adhesive tape is a viscoelastic material; therefore, the theory presented in Sec. 4.1. should be extended. For instance, the rate–temperature superposition principle could be adopted in this theory to describe the peel behavior of the structure with a viscoelastic adhesive tape [73,80].

5.2 Peel Phenomena for Flexible Packaging Heat Seal.

Food packaging is used to provide a barrier from harmful environmental factors such as microorganisms, light, and external gases; this barrier prevents or inhibits microbial and biochemical degradation and preserves inner aromatic gases to maintain flavor and freshness for the packed product. Therefore, the packaging should remain intact or sealed until consumer use. Flexible plastic packaging, which comprises thin film laminates, is a prevalent type of packaging that constitutes approximately 30% of the global packaging market and is increasing at an annual growth rate of 4.5%. This growth is primarily driven by the conversion of rigid packaging formats (e.g., glass, metal) to flexible packaging formats. Flexible packaging has several benefits over rigid packaging, such as being more convenient to carry and lighter in weight, which implies less energy is consumed for transportation. Flexible packaging is also more economical than rigid packaging [123].

Flexible packages are typically produced from materials like polymers, paper, and aluminum foil, utilizing a thin film or layered structure that is heat sealed [124]. This layering creates a laminate that combines unique properties and functionalities from different substances. The innermost layer is engineered for sealing, which is often referred to as the sealant layer [125]. Commonly, these sealant layers comprise thermally active materials such as thermoplastics, capable of adhering to not only similar materials but also to different substrates including plastic trays, bottles, metal cans, and glass containers. The process of sealing flexible packaging films can be accomplished through various techniques, with heat sealing being the most prevalent. Film packaging seals are primarily established via conductive heat, using a range of methods like seal bar, hot wire, hot air, pneumatic, ultrasound, and laser heating.

The seal performance is essential to ensure food safety and quality. In addition to strength, the barrier property is significant to inhibit food degradation and maintain freshness. It is noteworthy that heat-sealing applications frequently entail contact between two materials that are identical. The mechanisms contributing to adhesion and peel force involve the formation of molecular entanglements and co-crystallization at the interface between two polymer layers, among other factors, different from those discussed for release liners. These mechanisms are examined in more detail in the following sections.

5.2.1 Adhesion Mechanisms and Technical Approaches.

Heat sealing is a process that involves melting two seal sides of packaging materials at high temperatures, allowing interphase molecular chain diffusion and solidifying them at low temperatures to create a seal. It is essential to comprehend the adhesion mechanisms to devise specific strategies to enhance the adhesion strength. Based on an energy balance perspective (see Sec. 3), the adhesion mechanisms during the heat-sealing process can be classified into two main types: those related to the interface, characterized by the interfacial fracture energy (Γ), and those concerning bulk materials, for which the quantity dUd represents the energy dissipated within the bulk of the structure.

The main theoretical principles of forming a strong interface are to ensure close contact and robust interfacial bridging connections. The degree of contact depends largely on wettability, viscosity of sealant materials, surface contaminants, and so forth. Specifically, the wettability of the molten materials is the crucial parameter since it determines the ability to bond by filling gaps between two surfaces. The wetting behavior of a material is strongly related to its surface free energy (thermodynamic work of adhesion), and Dupre's equation is commonly used to estimate surface free energy (Eq. (36)).

In practice, some pressure is applied during heat sealing to improve adequate contact and wetting at the interface. Moreover, materials with lower viscosity at heating temperatures are preferable because they can fill gaps more easily in pouches or flow around contamination, thus achieving close contact and preventing leaks [126]. However, low-viscosity species may also reduce the overall bonding strength due to reduced entanglement. Therefore, polymer blends—for example, different grades of polyethylene—are also used to tune the heat seal strength. A similar strategy is applied when considering chain diffusion (see below), as the viscosity is closely related to chain diffusion.

Diffusion and entanglements determine the strength and effectiveness of the interfacial connections between the polymer layers. However, depending on the specific characteristics of the polymers involved, other mechanisms may also play a significant role. For example, for semicrystalline polymers, the crystallization process affects the number and the size of the crystals that form at the interface [127]. For polar polymers such as polylactic acid (PLA) and ethylene vinyl acetate, intermolecular interactions due to hydrogen bonds, polarity, and ionic groups are also important.

Since diffusion and entanglements are universal mechanisms, the following sections provide more detailed discussions of each. The investigation of chain diffusion requires a thorough understanding of the molecular level mobility of the material, which is intrinsically linked to its thermal attributes. For example, thermoplastics must be heated to enable diffusion and seal. Therefore, the thermal properties of sealant materials are of high importance. For amorphous polymers, the glass transition temperature (Tg) is most important as molecular segments transition from a glassy to a rubbery state with significant chain movement. For crystalline or semicrystalline materials, the melting temperature (Tm) is important as the crystalline phases begin to melt, and the material starts to flow. For instance, the seal initiation of semicrystalline low-density polyethylene (LDPE) or linear low-density polyethylene (LLDPE) occurs when the amorphous fraction increases by heating as crystalline regions melt and polymer chains become more mobile.

It is noteworthy that process parameters, such as interfacial temperature and seal time, are critical for sufficient diffusion and entanglements [128]. With increasing temperature, diffusion and entanglement are enhanced. The temperature at the interface should ensure the melted state of the sealant materials since melted sealant materials can facilitate better interface chain diffusion. A rapid increase in the seal strength is observed with increasing temperature until a plateau value is reached around Tm. Moreover, the sealing temperature needs to be controlled to be lower than the decomposition temperature (Td).

The diffusion rates (chain mobility) of molecule chains depend primarily on the compatibility of the two sealant materials, molecular weight, chain branching, and so forth. For example, polymer chains with lower molecular weight tend to diffuse more rapidly. However, to obtain superior bonding strength, chains of higher molecular weight polymers are required to increase the potential for entanglements [129]—a characteristic that enhances the interfacial fracture energy and thus has a crucial effect on the final seal strength. The process of forming entanglements begins from the initial stages of sealing and continues through the cooling process. A more comprehensive discussion on fracture energy can be found in Sec. 3.4. In addition, there is a viscosity trade-off for sealant materials: slower diffusion of longer chains implies that it takes longer for these chains to form entanglements. Therefore, exploring the interplay between molecular weight, diffusion time, and fracture energy could be a compelling avenue for future research. In practice, the use of a combination of various molecular weight materials can thus balance the rate of the sealing process and the adhesion strength.

The theoretical mechanisms that underlie the behavior of bulk materials involve modifying the energy terms that originate from the bulk materials. For example, the incorporation of additives can significantly alter the properties of bulk materials, thereby affecting the energy dissipation from bulk materials. Furthermore, additives and processing aids are organic or inorganic molecules that are introduced in small quantities to the polymer matrix to adjust the properties of packaging films and/or enhance processing. According to Ref. [130], some common examples of stabilizers, modifiers, and other additives are pigments, slip agents, antiblocking, antistatic, process aids, nucleating agents, tackifiers, tougheners, and fillers.

5.2.2 Characterization of Adhesion Performance.

Several ASTM standards are used to assess seal performance. Among them, ASTM F88/F88M-09 is widely used [131]. Seal strength is a measurement of the forces required to separate two sealed films. In this method, film samples are cut into strips of a specified size (e.g., 1 × 6 in.) using a die cutter of this dimension. A T-peel test is then conducted on the strips using an Instron Universal Testing Machine. T-peel tests are performed with an appropriate load cell (e.g., 5-kg load cell) at a predetermined separation rate (e.g., 10 cm/min). The maximum and average peel forces are recorded as the seal/peel strength.

Note that seal separation may also involve potential deformation or elongation of the seal films during the test. The influence of the elongation of the tape can be evaluated through energy-based and stress-based analyses. Specifically, for the elastic tape, the influence of elongation deformation can be calculated by Eq. (4).

To minimize the impact of the film material deformation, a piece of masking tape is often applied on the back of the test film strip to prevent stretching of the film. This masking tape allows for testing of adhesion between two sealed layers rather than the strength of the film itself. There are also various types of nondestructive testing and evaluation techniques that can potentially be used for the inspection of heat-sealing regions [132].

5.2.3 Market, Prospect, and Opportunity.

Polyolefins, especially LDPE or LLDPE, constitute the most prevalent sealing materials. Other thermoplastics such as ethylene vinyl acetate and ionomers, polyesters, including poly(ethylene terephthalate), and poly(butylene succinate) are also employed. In response to the growing demand for more environmentally friendly packaging solutions in the food packaging industry, changes encompass innovations in structure designs that facilitate better recyclability as well as bio-based and/or biodegradable alternatives such as cellulose-, starch-, PLA-based films. Further research is anticipated to design the material properties of such biodegradable alternatives and optimize their seal performance.

The selection and optimization of new materials for flexible packaging can be greatly facilitated and differentiated by applying a predictive engineering approach. Some studies have presented models of the influence of heat-sealing process parameters, such as temperature, dwell time, and pressure, on the peel strength [133,134]. However, there is a lack of research on the evaluation and modeling of peel phenomena during the peeling process. Therefore, there are opportunities to model the behavior of flexible packaging in a more comprehensive way. For instance, a parametric finite element model could be developed to simulate the progressive failure of the heat seal interface in a typical T-peel test. The seal interface could be modeled with a cohesive zone method.

To conclude, the seal strength is influenced by various factors such as the materials used, process conditions, and packaging designs. Peel modeling can facilitate a deeper understanding of peel failure modes and enable peel strength optimization. Predictive modeling for flexible packaging may contribute to the improvement and optimization of the material selection of new sustainable polymers to achieve seals with the required barrier and bonding strength.

5.3 Peeling Soft Adhesives From Human Skin.

A final commonly used application of soft adhesives on soft substrates encompasses medical adhesives, specifically those used on skin for securing wound dressings or for securing equipment like monitoring and life support devices, such as surgical drapes, steric-strips, electrodes, antimicrobial drape, wound dressings, and bandages. Adhesives for use in general biomedical applications are a rapidly growing industry with a market share of nearly $9 billion in 2021 [135]. By 2026, this industry is expected to reach nearly $12 billion, exhibiting double-digit annual growth rates [135137]. This rapid growth is attributed to several factors, including their increased use in medical devices, implants, dental applications, and as replacements for surgical sutures or stapling to improve patient outcomes [135,137,138].

There are numerous published reviews specifically on skin adhesives, which can be largely divided into two categories: those focused on technical and materials approaches to control adhesion [139142] and those focused on clinical cases by physicians [143145]. To avoid repeating information already contained in these works and others, in this section, we provide only a brief overview of these works, general design criteria for skin adhesives, and unique approaches for characterizing their release performance. Several types of adhesives are used in applications for skin contact. Major categories include acrylic adhesives, silicone adhesives, polyurethane adhesives, and rubber-based adhesives, among others. Each type of adhesive has advantages and disadvantages, summarized in Table 2. Other categories for medical tapes and adhesives include polyolefins, hydrocolloids, and bio-based materials and hydrogels [140,145147]. In particular, bio-based adhesives for such biomedical applications are a rapidly growing industry, comprising an over $2 billion market in the United States in 2021 that is expected to have a compound annual growth rate of 14% between 2022 and 2027 [147,148].

Table 2

Typical PSAs for skin applications

Acrylic adhesivesSilicone adhesivesPolyurethane adhesivesRubber-based adhesives
Adhesion strengthMedium to highLow to mediumMedium to highMedium to high
TackinessLow to highLow to mediumLow to mediumTypically high
Cohesion strengthLow to highMedium to highLow to mediumMedium to high
Environmental resistanceModerateExcellentHighModerate to high
FlexibilityModerate to HighExcellentHighModerate to High
Potential for skin irritationLow to moderateLowLow to moderateLow
DurabilityGoodHighHighModerate to high
CostModerateHighModerate to highLow to moderate
Acrylic adhesivesSilicone adhesivesPolyurethane adhesivesRubber-based adhesives
Adhesion strengthMedium to highLow to mediumMedium to highMedium to high
TackinessLow to highLow to mediumLow to mediumTypically high
Cohesion strengthLow to highMedium to highLow to mediumMedium to high
Environmental resistanceModerateExcellentHighModerate to high
FlexibilityModerate to HighExcellentHighModerate to High
Potential for skin irritationLow to moderateLowLow to moderateLow
DurabilityGoodHighHighModerate to high
CostModerateHighModerate to highLow to moderate

Note: Data are summarized from Refs. [142,146,149].

Compared with other PSA applications, the skin presents a unique and challenging adherend, distinct from other substrates due to its distinctive properties. The skin is a soft substrate with a low modulus, relatively low surface energy and a textured surface [146,150]. Additionally, the skin is a dynamic substrate undergoing rapid regeneration; as such, skin adhesives must adhere to a substrate that is constantly being shed. The skin is also susceptible to contamination from substances such as sweat and grease, which alter the surface energy and serves as a weak boundary layer, leading to poor long-term performance. Finally, the skin is prone to allergies caused by many chemicals, all of which contribute to the complexity of developing new adhesives.

Despite the challenges associated with the skin, the development of new adhesives with enhanced performance, such as controlled adhesion, hypoallergenic properties, and gentle removal, along with improved sustainability, is crucial to meet robust market demands, especially in medical and wearable electronics applications. Similar to other PSA applications, adhesives designed for skin applications prioritize three key performance factors: peel adhesion strength (resistance to removal by peeling), cohesion strength (resistance to shearing forces), and tack strength (speed of adherence to the adherent). However, skin adhesives must also have good water resistance and must pass strict quality control standards. Additional performance metrics are critical for designing these systems, including biocompatibility, skin permeability, skin trauma control, drug delivery, and others [137,139,141,145]. For example, traditional acrylic adhesives have low to moderate costs and medium to high adhesion strengths (Table 2), but cause more deformation and trauma to the skin upon removal [142,145,146,149,151]. Thus, to develop a successful product, a fundamental understanding of the peeling mechanism of the adhesive from the skin is crucial.

Beyond assessing performance via standard tests of peel adhesion strength, cohesion strength, and tack strength, several other tests have been used to evaluate clinical outcomes associated with the use of skin adhesives. Due to the wide range of skin adhesive functions and properties, the appropriate in vitro or in vivo tests to evaluate performance are application specific. For example, clinical studies on skin adhesive tapes have performed peel tests on human patients and evaluated transepidermal water loss, erythema (skin redness), and hydration following adhesive removal [145,152]. Additionally, several studies have evaluated skin adhesives in vitro using perspiration simulators to better understand the adhesive performance during sweating [153155]. Clinical works have also measured alterations in skin barrier function, skin tearing and stripping, and skin dermatitis following adhesive removal [145,151,152]. In vivo tests of topical skin adhesive (TSA) performance also include measures of wound bursting strength and TSA flexibility [145,156158]. As this wide range of application-specific testing protocols demonstrates, newly developed skin and tissue adhesives require additional performance testing versus standard PSAs to account for biological damage, prior to their adoption by the medical industry.

6 Conclusion and Prospect

This review critically examines the theoretical foundations of peeling mechanics, with a special focus on peeling from soft substrates. It emphasizes the dual perspectives of energy and stress analysis. Additionally, the article discusses three potential industrial applications of peel mechanics. This work highlights the critical roles of viscoelastic dissipation and deformation behaviors in influencing peeling mechanics, particularly within the context of PSA-release liner systems and flexible packaging heat seal systems.

For the PSA-release liner system, the application of the stress-based analysis from Sec. 4.1 illuminates potential pathways for a more nuanced understanding of peel mechanics. This framework, albeit simplified to treat the tape as a homogeneous elastic material, hints at the complex interplay between material properties and peel behavior. In real-world scenarios, the distinct viscoelastic nature of the adhesive material emerges as a pivotal factor influencing peel behavior. The insights from Sec. 3.1.3 on evaluating viscoelastic dissipation energy further enrich our understanding of these systems and suggest a need to extend the theoretical models to encapsulate the viscoelastic characteristics of adhesive tapes. Such adaptations could leverage the rate–temperature superposition principle to yield a more accurate depiction of peel behavior under diverse conditions.

For the flexible packaging heat seal system, the theoretical concepts discussed in Sec. 3 offer a foundational understanding of the adhesion mechanisms pivotal during the heat-sealing process. The delineation of adhesion mechanisms into interface-related and bulk material-related categories provides a framework for assessing seal strength. Furthermore, the discussion on polymer chain characteristics and their diffusion rates underscores the nuanced balance between molecular weight and bonding strength, highlighting the essential role of polymer entanglement in augmenting interfacial fracture energy and, by extension, the durability of the seal.

This review also highlights several potential research avenues that remain untapped, evidenced by the unoccupied cells in Table 1. For instance, adhesive structures with both linear viscoelastic tape and linear viscoelastic substrates have yet to be extensively examined. Further exploration into these topics is not only academically stimulating but also holds substantial practical implications. For example, considering that both soft tissues and adhesives exhibit energy dissipation characteristics, developing a model that accounts for the viscoelastic properties of both the tape and the substrate could significantly advance bioadhesive applications. These gaps in research may stem from numerous factors such as the complexity of the problems, which renders mathematical models challenging to develop or apply. Often, to achieve a qualitative understanding of the peeling process, many mathematical theories may simplify the scenario by focusing on linear elastic materials, thereby neglecting more intricate physical phenomena like viscoelastic dissipation. Although these simplifications enable easier mathematical handling, they can omit critical physical insights. Nevertheless, these uncharted territories can present promising avenues for future research, holding the potential to further expand and deepen our understanding in this domain.

Acknowledgment

Y. X., C. J. E., L. F. F., and M. A. C. thank The Dow Chemical Company for supporting this collaborative research. The authors thank Tim Mitchell and Michaeleen Pacholski for useful discussions.

Conflict of Interest

There are no conflicts of interest.

Data Availability Statement

The authors attest that all data for this study are included in the paper.

Appendix A: The Bending of Linear Elastic Tape

For linear elastic material, EIdϕ/ds=M(s) holds, Then Eq. (6) can be reorganized as follows [66]:
(A1)
By integrating Eq. (A1), we obtain the following:
(A2)
By substituting (EIdφds)A2=MA2=(Fm)2 into Eq. (A2), the following equation holds:
(A3)
When θ00, Eq. (A3) is simplified as follows:
(A4)

Appendix B: Momentum Balance for Tape With a Circular Shape

Derail et al. [75,76] developed a theoretical method based on the balance of angular momentum to capture the peeling process of the adhesive structure with an elastic/elastic–plastic tape and viscoelastic adhesive.

Applying the balance of angular momentum of the whole peeling arm at point O, the following equation holds:
(B1)
where Rc is the radius of curvature, assumed to be a constant; σp is the stress in the peeling arm at the point O, where the deformation of the adhesive vanishes; and σa is the stress in the adhesive (see Fig. 11). They can be evaluated by using the corresponding constitutive equations.
Fig. 11
Illustration of the structure with soft adhesive with the assumption that the radius of curvature of the peeling arm is a constant Rc. An element of the peeling arm at the point O′ is magnified.
Fig. 11
Illustration of the structure with soft adhesive with the assumption that the radius of curvature of the peeling arm is a constant Rc. An element of the peeling arm at the point O′ is magnified.
Close modal

The Stress of Peeling the Arm.

When the peeling arm is elastic, the stress can be calculated as follows [75]:
(B2)
Then, the first term in Eq. (B1) can be rewritten as follows:
(B3)
if the peeling arm undergoes plastic deformation [75]. Then, the first term Eq. (B1) can be rewritten as follows:
(B4)
where σp0 is the plastic yield stress of the tape and lp is the plastically deformed thickness (see the inset of Fig. 11).

The Stress of the Soft Adhesive.

To evaluate the stress in the soft adhesive (the second term of Eq. (B1)), a viscoelastic model is required. There are many available viscoelastic models [160,161]. Derail et al. [75,76] assume that the adhesive experiences an elongational deformation and used the Kaye and Bernstein-Kearsley-Zapas (K-BKZ) model [162] to capture its mechanical behavior:
(B5)
where m(t)=dG(t)/dt0 is the memory function, G(t) is the relaxation modulus that can be measured by rheology, C1 is the finger strain tensor (the inverse of the right Cauchy–Green deformation tensor), λ is the stretch ratio, and hd(λ) is the damping function that corrects the nonlinearity in the relaxation modulus [163]. Derail et al. [75,76] adopted the damping function from the study by Wagner [164] as follows:
(B6)
The normal strain e in the adhesive can be expressed as follows:
(B7)
The stretch ratio of the adhesive during the peeling process can be given as follows:
(B8)
The angular position θ can be calculated as follows:
(B9)
where V is the peeling velocity. Combining Eqs. (B5), (B8), and (B9), the second term of Eq. (B1) can be evaluated. The only remaining term is the upper limit θM, which could be determined from the fracture criterion. We will discuss this in the following subsection.

Determination of the Angular Position of the Failure Locus θM.

The location of the failure locus θM depends on the failure mode. When cohesive failure occurs, the cohesive break at the constant Hencky strain ε=ln(1+e)=4.5 in uniaxial extension deformation [75]. The locus θM can then be calculated from Eq. (B7) as follows:
(B10)
To capture the interfacial adhesive failure, Ref. [76] applied fracture criterion stemming from the trumpet model [52] as follows:
(B11)
where Cf is a parameter that needs to be fitted from experiments, and u is the crack (peel front) opening, which is calculated as follows:
(B12)

For a given peeling velocity, the value of θM can be evaluated. The two criteria (Eqs. (B10) and (B12)) coexist. The cohesive failure transitions to interfacial adhesive failure when θM evaluated from Eq. (B10) is smaller than θM evaluated from Eq. (B12).

Remarks on This Angular Momentum-Based Approach.

Although this angular momentum-based approach is developed for the elastic (Eq. (B3)/elastic–plastic (Eq. B4)) tape and the nonlinear viscoelastic adhesive (Eq. (B5)), this approach could be extended for considering other materials systems. The only requirement is to replace the corresponding constitutive model with the necessary constitutive equations. For instance, the viscoelastic model can be utilized in the first term of Eq. (B1) to capture the rate dependence of the tape. Note that in this model, the peeling arm is assumed to be circular, which may not be correct in the real application. The shape of the peeling arm can be explicitly evaluated using the bending equations, a topic elaborated upon in Sec. 4.3.

References

1.
Kendall
,
K.
,
1971
, “
The Adhesion and Surface Energy of Elastic Solids
,”
J. Phys. D: Appl. Phys.
,
4
(
8
), pp.
1186
1195
.
2.
Pesika
,
N. S.
,
Tian
,
Y.
,
Zhao
,
B. X.
,
Rosenberg
,
K.
,
Zeng
,
H. B.
,
McGuiggan
,
M. P.
,
Autumn
,
K.
, and
Israelachvili
,
J.
,
2007
, “
Peel-Zone Model of Tape Peeling Based on the Gecko Adhesive System
,”
J. Adhes.
,
83
(
4
), pp.
383
401
.
3.
Rivlin
,
R. S.
,
1944
, “
The Effective Work of Adhesion
,”
Paint Tech.
,
9
(
106
), pp.
215
218
.
4.
Moidu
,
A. K.
,
Sinclair
,
A. N.
, and
Spelt
,
J. K.
,
1995
, “
Analysis of the Peel Test—Prediction of Adherend Plastic Dissipation and Extraction of Fracture Energy in Metal-to-Metal Adhesive Joints
,”
J. Test. Eval.
,
23
(
4
), pp.
241
253
.
5.
Moidu
,
A. K.
,
Sinclair
,
A. N.
, and
Spelt
,
J. K.
,
1998
, “
On the Determination of Fracture Energy Using the Peel Test
,”
J. Test. Eval.
,
26
(
3
), pp.
247
254
.
6.
Chen
,
B.
,
Wu
,
P.
, and
Gao
,
H.
,
2009
, “
Pre-tension Generates Strongly Reversible Adhesion of a Spatula Pad on Substrate
,”
J. R. Soc. Interface
,
6
(
35
), pp.
529
537
.
7.
Tian
,
Y.
,
Pesika
,
N.
,
Zeng
,
H.
,
Rosenberg
,
K.
,
Zhao
,
B.
,
McGuiggan
,
P.
,
Autumn
,
K.
, and
Israelachvili
,
J.
,
2006
, “
Adhesion and Friction in Gecko Toe Attachment and Detachment
,”
Proc. Natl. Acad. Sci. USA
,
103
(
51
), pp.
19320
19325
.
8.
Ceglie
,
M.
,
Menga
,
N.
, and
Carbone
,
G.
,
2022
, “
The Role of Interfacial Friction on the Peeling of Thin Viscoelastic Tapes
,”
J. Mech. Phys. Solids
,
159
, pp.
1
9
.
9.
Thouless
,
M. D.
, and
Yang
,
Q. D.
,
2008
, “
A Parametric Study of the Peel Test
,”
Int. J. Adhes. Adhes.
,
28
(
4–5
), pp.
176
184
.
10.
Sauer
,
R. A.
,
2011
, “
The Peeling Behavior of Thin Films With Finite Bending Stiffness and the Implications on Gecko Adhesion
,”
J. Adhes.
,
87
(
7–8
), pp.
624
643
.
11.
12.
Eley
,
R. R.
, and
Petrash
,
S.
,
2006
, “Surface Energetics of Polymers and Rubbers, Characterization of,”
Encyclopedia of Analytical Chemistry
,
John Wiley & Sons, Ltd
,
Hoboken, NJ
, pp.
1
15
.
13.
Sun
,
S.
,
Li
,
M.
, and
Liu
,
A.
,
2013
, “
A Review on Mechanical Properties of Pressure Sensitive Adhesives
,”
Int. J. Adhes. Adhes.
,
41
(
1
), pp.
98
106
.
14.
Park
,
K.
, and
Paulino
,
G. H.
,
2011
, “
Cohesive Zone Models: A Critical Review of Traction-Separation Relationships Across Fracture Surfaces
,”
ASME Appl. Mech. Rev.
,
64
(
6
), p.
060802
.
15.
Needleman
,
A.
,
2014
, “
Some Issues in Cohesive Surface Modeling
,”
Mechanics for the World: Proceedings of the 23rd International Congress of Theoretical and Applied Mechanics, ICTAM2012, 10
,
Beijing, China
,
Aug. 19–24, 2012
, pp.
221
246
.
16.
Wciślik
,
W.
, and
Pała
,
T.
,
2021
, “
Selected Aspects of Cohesive Zone Modeling in Fracture Mechanics
,”
Metals
,
11
(
2
), p.
302
.
17.
Sauer
,
R. A.
,
2015
, “
A Survey of Computational Models for Adhesion
,”
J. Adhes.
,
92
(
2
), pp.
81
120
.
18.
Gu
,
Z.
,
Li
,
S.
,
Zhang
,
F.
, and
Wang
,
S.
,
2016
, “
Understanding Surface Adhesion in Nature: A Peeling Model
,”
Adv. Sci.
,
3
(
7
), p.
1500327
.
19.
Bartlett
,
M. D.
,
Case
,
S. W.
,
Kinloch
,
A. J.
, and
Dillard
,
D. A.
,
2023
, “
Peel Tests for Quantifying Adhesion and Toughness: A Review
,”
Prog. Mater. Sci.
,
137
, pp.
1
133
.
20.
Kinloch
,
A. J.
,
Lau
,
C. C.
, and
Williams
,
J. G.
,
1994
, “
The Peeling of Flexible Laminates
,”
Int. J. Fract.
,
66
(
1
), pp.
45
70
.
21.
Suo
,
Z.
, and
Hutchinson
,
J. W.
,
1990
, “
Interface Crack Between Two Elastic Layers
,”
Int. J. Fract.
,
43
(
1
), pp.
1
18
.
22.
Hutchinson
,
J. W.
, and
Suo
,
Z.
,
1991
, “
Mixed Mode Cracking in Layered Materials
,”
Adv. Appl. Mech.
,
29
(
1
), pp.
63
191
.
23.
Thouless
,
M. D.
, and
Jensen
,
H. M.
,
1992
, “
Elastic Fracture-Mechanics of the Peel-Test Geometry
,”
J. Adhes.
,
38
(
3–4
), pp.
185
197
.
24.
Lu
,
Z. X.
,
Yu
,
S. W.
,
Wang
,
X. Y.
, and
Feng
,
X. Q.
,
2007
, “
Effect of Interfacial Slippage in Peel Test: Theoretical Model
,”
Eur. Phys. J. E
,
23
(
1
), pp.
67
76
.
25.
Kaelble
,
D. H.
,
1959
, “
Theory and Analysis of Peel Adhesion—Mechanisms and Mechanics
,”
Trans. Soc. Rheol.
,
3
(
1
), pp.
161
180
.
26.
Ghatak
,
A.
,
2010
, “
Peeling Off an Adhesive Layer With Spatially Varying Modulus
,”
Phys. Rev. E
,
81
(
2
), p.
021603
.
27.
Kendall
,
K.
,
1975
, “
Thin-Film Peeling—The Elastic Term
,”
J. Phys. D: Appl. Phys.
,
8
(
13
), pp.
1449
1452
.
28.
Kendall
,
K.
,
1975
, “
Transition Between Cohesive and Interfacial Failure in a Laminate
,”
Proc. R. Soc. London Ser., A
,
344
(
1637
), pp.
287
302
.
29.
Anderson
,
G. P.
,
DeVries
,
K. L.
, and
Williams
,
M. L.
,
1976
, “
The Peel Test in Experimental Adhesive-Fracture Mechanics
,”
Exp. Mech.
,
16
(
1
), pp.
11
15
.
30.
Kim
,
K. S.
, and
Aravas
,
N.
,
1988
, “
Elastoplastic Analysis of the Peel Test
,”
Int. J. Solids Struct.
,
24
(
4
), pp.
417
435
.
31.
Tsai
,
K.-H.
, and
Kim
,
K.-S.
,
1993
, “
Stick-Slip in the Thin Film Peel Test—I. The 90 Peel Test
,”
Int. J. Solids Struct.
,
30
(
13
), pp.
1789
1806
.
32.
He
,
L.
,
Lou
,
J.
,
Kitipornchai
,
S.
,
Yang
,
J.
, and
Du
,
J.
,
2019
, “
Peeling Mechanics of Hyperelastic Beams: Bending Effect
,”
Int. J. Solids Struct.
,
167
(
1
), pp.
184
191
.
33.
Chen
,
H.
,
Feng
,
X.
,
Huang
,
Y.
,
Huang
,
Y.
, and
Rogers
,
J. A.
,
2013
, “
Experiments and Viscoelastic Analysis of Peel Test With Patterned Strips for Applications to Transfer Printing
,”
J. Mech. Phys. Solids
,
61
(
8
), pp.
1737
1752
.
34.
Pierro
,
E.
,
Afferrante
,
L.
, and
Carbone
,
G.
,
2020
, “
On the Peeling of Elastic Tapes From Viscoelastic Substrates: Designing Materials for Ultratough Peeling
,”
Tribol. Int.
,
146
(
9
), p.
106060
.
35.
Peng
,
Z.
, and
Chen
,
S.
,
2015
, “
Effect of Bending Stiffness on the Peeling Behavior of an Elastic Thin Film on a Rigid Substrate
,”
Phys. Rev. E
,
91
(
4
), p.
042401
.
36.
Williams
,
J. G.
,
Hadavinia
,
H.
, and
Cotterell
,
B.
,
2005
, “
Anisotropic Elastic and Elastic-Plastic Bending Solutions for Edge Constrained Beams
,”
Int. J. Solids Struct.
,
42
(
18–19
), pp.
4927
4946
.
37.
Kawashita
,
L. F.
,
2006
, “
The Peeling of Adhesive Joints
,”
Ph.D. thesis
,
University of London
,
London
.
39.
Kinloch
,
A.
, and
Williams
,
J.
,
2002
, “
The Mechanics of Peel Tests
,”
Adhes. Sci. Eng.
,
1
(
8
), pp.
273
301
.
40.
Zhang Newby
,
B.
, and
Chaudhury
,
M. K.
,
1997
, “
Effect of Interfacial Slippage on Viscoelastic Adhesion
,”
Langmuir
,
13
(
6
), pp.
1805
1809
.
41.
Ghatak
,
A.
,
Vorvolakos
,
K.
,
She
,
H.
,
Malotky
,
D. L.
, and
Chaudhury
,
M. K.
,
2000
, “
Interfacial Rate Processes in Adhesion and Friction
,”
J. Phys. Chem. B
,
104
(
17
), pp.
4018
4030
.
42.
Raphael
,
E.
, and
De Gennes
,
P. G.
,
1992
, “
Rubber-Rubber Adhesion With Connector Molecules
,”
J. Phys. Chem.
,
96
(
10
), pp.
4002
4007
.
43.
Brochard-Wyart
,
F.
,
De Gennes
,
P. G.
,
Leger
,
L.
,
Marciano
,
Y.
, and
Raphael
,
E.
,
1994
, “
Adhesion Promoters
,”
J. Phys. Chem.
,
98
(
38
), pp.
9405
9410
.
44.
Creton
,
C.
,
Brown
,
H. R.
, and
Shull
,
K. R.
,
1994
, “
Molecular Weight Effects in Chain Pullout
,”
Macromolecules
,
27
(
12
), pp.
3174
3183
.
45.
Gaire
,
B.
,
Wilson
,
M. C.
,
Singla
,
S.
, and
Dhinojwala
,
A.
,
2022
, “
Connection Between Molecular Interactions and Mechanical Work of Adhesion
,”
ACS Macro Lett.
,
11
(
11
), pp.
1285
1290
.
46.
Owens
,
D. K.
, and
Wendt
,
R. C.
,
1969
, “
Estimation of the Surface Free Energy of Polymers
,”
J. Appl. Polym. Sci.
,
13
(
8
), pp.
1741
1747
.
47.
Gent
,
A. N.
, and
Schultz
,
J.
,
1972
, “
Effect of Wetting Liquids on the Strength of Adhesion of Viscoelastic Material
,”
J. Adhes.
,
3
(
4
), pp.
281
294
.
48.
Andrews
,
E.
, and
Kinloch
,
A. J.
,
1973
, “
Mechanics of Adhesive Failure. II
,”
Proc. R. Soc. London. A
,
332
(
1590
), pp.
401
414
.
49.
Benyahia
,
L.
,
Verdier
,
C.
, and
Piau
,
J.-M.
,
1997
, “
The Mechanisms of Peeling of Uncross-Linked Pressure Sensitive Adhesives
,”
J. Adhes.
,
62
(
1–4
), pp.
45
73
.
50.
Du
,
J.
,
Lindeman
,
D. D.
, and
Yarusso
,
D. J.
,
2004
, “
Modeling the Peel Performance of Pressure-Sensitive Adhesives
,”
J. Adhes.
,
80
(
7
), pp.
601
612
.
51.
Saulnier
,
F.
,
Ondarcuhu
,
T.
,
Aradian
,
A.
, and
Raphael
,
E.
,
2004
, “
Adhesion Between a Viscoelastic Material and a Solid Surface
,”
Macromolecules
,
37
(
3
), pp.
1067
1075
.
52.
de Gennes
,
P. G.
,
1996
, “
Soft Adhesives
,”
Langmuir
,
12
(
19
), pp.
4497
4500
.
53.
Maugis
,
D.
, and
Barquins
,
M.
,
1978
, “
Fracture Mechanics and the Adherence of Viscoelastic Bodies
,”
J. Phys. D: Appl. Phys.
,
11
(
14
), pp.
1989
2023
.
54.
Peng
,
Z. L.
,
Wang
,
C.
,
Chen
,
L.
, and
Chen
,
S. H.
,
2014
, “
Peeling Behavior of a Viscoelastic Thin-Film on a Rigid Substrate
,”
Int. J. Solids Struct.
,
51
(
1
), pp.
4596
4603
.
55.
Zhu
,
Z. M.
,
Yang
,
Z. R.
,
Xia
,
Y.
, and
Jiang
,
H.
,
2021
, “
Controllable Peeling of an Elastic Strip on a Viscoelastic Substrate
,”
Eng. Fract. Mech.
,
256
, pp.
1
13
.
56.
Chaudhury
,
M. K.
,
1999
, “
Rate-Dependent Fracture at Adhesive Interface
,”
J. Phys. Chem. B
,
103
(
31
), pp.
6562
6566
.
57.
Song
,
Z.
,
Shen
,
T.
,
Vernerey
,
F. J.
, and
Cai
,
S.
,
2021
, “
Force-Dependent Bond Dissociation Explains the Rate-Dependent Fracture of Vitrimers
,”
Soft Matter
,
1
(
27
), pp.
6669
6674
.
58.
Afferrante
,
L.
, and
Carbone
,
G.
,
2016
, “
The Ultratough Peeling of Elastic Tapes From Viscoelastic Substrates
,”
J. Mech. Phys. Solids
,
96
(
3
), pp.
223
234
.
59.
Menga
,
N.
,
Afferrante
,
L.
,
Pugno
,
N. M.
, and
Carbone
,
G.
,
2018
, “
The Multiple V-Shaped Double Peeling of Elastic Thin Films From Elastic Soft Substrates
,”
J. Mech. Phys. Solids
,
113
(
1
), pp.
56
64
.
60.
Ciavarella
,
M.
,
McMeeking
,
R. M.
, and
Cricrì
,
G.
,
2021
, “
On the Afferrante-Carbone Theory of Ultratough Tape Peeling
,”
FU Mech. Eng.
,
21
(
4
), pp.
727
735
.
61.
Carbone
,
G.
, and
Mangialardi
,
L.
,
2008
, “
Analysis of the Adhesive Contact of Confined Layers by Using a Green's Function Approach
,”
J. Mech. Phys. Solids
,
56
(
2
), pp.
684
706
.
62.
Perrin
,
H.
,
Eddi
,
A.
,
Karpitschka
,
S.
,
Snoeijer
,
J. H.
, and
Andreotti
,
B.
,
2019
, “
Peeling an Elastic Film From a Soft Viscoelastic Adhesive: Experiments and Scaling Laws
,”
Soft Matter
,
15
(
4
), pp.
770
778
.
63.
Lavoie
,
S. R.
,
Long
,
R.
, and
Tang
,
T.
,
2015
, “
Rate Dependent Fracture of a Double Cantilever Beam With Combined Bulk and Interfacial Dissipation
,”
Int. J. Solids Struct.
,
75–76
(
1
), pp.
277
286
.
64.
Spies
,
G.
,
1953
, “
The Peeling Test on Redux-Bonded Joints: A Theoretical Analysis of the Test Devised by Aero Research Limited
,”
Aircraft Eng. Aerosp. Technol
,
25
(
3
), pp.
64
70
.
65.
Bikerman
,
J. J.
,
1957
, “
Theory of Peeling Through a Hookean Solid
,”
ASME J. Appl. Phys.
,
28
(
12
), pp.
1484
1485
.
66.
Kaelble
,
D. H.
,
1960
, “
Theory and Analysis of Peel Adhesion—Bond Stresses and Distributions
,”
Trans. Soc. Rheol.
,
4
(
1
), pp.
45
73
.
67.
Dillard
,
D. A.
,
1989
, “
Bending of Plates on Thin Elastomeric Foundations
,”
ASME J. Appl. Mech.
,
56
(
2
), pp.
382
386
.
68.
Plaut
,
R. H.
,
Hwang
,
D.
,
Lee
,
C.
,
Bartlett
,
M. D.
, and
Dillard
,
D. A.
,
2022
, “
Peeling of Finite-Length Elastica on Winkler Foundation Until Complete Detachment
,”
Int. J. Solids Struct.
,
256
, pp.
1
14
.
69.
Plaut
,
R. H.
, and
Dillard
,
D.
,
2023
, “
Peeling of Finite-Length Plates From an Elastomeric Foundation: A 1-D Cylindrical Bending Solution
,”
ASME J. Appl. Mech.
,
90
(
9
), p.
091001
.
70.
Ghatak
,
A.
,
Mahadevan
,
L.
, and
Chaudhury
,
M. K.
,
2005
, “
Measuring the Work of Adhesion Between a Soft Confined Film and a Flexible Plate
,”
Langmuir
,
21
(
4
), pp.
1277
1281
.
71.
Ghatak
,
A.
,
2011
, “
Mechanics of Peeling of a Flexible Adherent Off a Thin Layer of Adhesive
,”
Mechanics over Micro and Nano Scales
,
Springer
,
New York
, pp.
171
191
.
72.
Mukherjee
,
B.
,
Dillard
,
D. A.
,
Moore
,
R. B.
, and
Batra
,
R. C.
,
2016
, “
Debonding of Confined Elastomeric Layer Using Cohesive Zone Model
,”
Int. J. Adhes. Adhes.
,
66
(
1
), pp.
114
127
.
73.
Kaelble
,
D. H.
,
1964
, “
Theory and Analysis of Peel Adhesion: Rate-Temperature Dependence of Viscoelastic Interlayers
,”
J. Colloid Sci.
,
19
(
5
), pp.
413
424
.
74.
Kaelble
,
D. H.
,
1965
, “
Peel Adhesion: Micro-Fracture Mechanics of Interfacial Unbonding of Polymers
,”
Trans. Soc. Rheol.
,
9
(
2
), pp.
135
163
.
75.
Derail
,
C.
,
Allal
,
A.
,
Marin
,
G.
, and
Tordjeman
,
P.
,
1997
, “
Relationship Between Viscoelastic and Peeling Properties of Model Adhesives. Part 1. Cohesive Fracture
,”
J. Adhes.
,
61
(
1–4
), pp.
123
157
.
76.
Derail
,
C.
,
Allal
,
A.
,
Marin
,
G.
, and
Tordjeman
,
P.
,
1998
, “
Relationship Between Viscoelastic and Peeling Properties of Model Adhesives. Part 2. The Interfacial Fracture Domains
,”
J. Adhes
,
68
(
3–4
), pp.
203
228
.
77.
Li
,
Y.
,
Pesika
,
N. S.
,
Zhou
,
M.
, and
Tian
,
Y.
,
2018
, “
Spring Contact Model of Tape Peeling: A Combination of the Peel-Zone Approach and the Kendall Approach
,”
Front. Mech. Eng.
,
4
(
22
), pp.
1
11
.
78.
Gent
,
A. N.
, and
Hamed
,
G. R.
,
1975
, “
Peel Mechanics
,”
J. Adhes.
,
7
(
2
), pp.
91
95
.
79.
Gent
,
A. N.
, and
Hamed
,
G. R.
,
1977
, “
Peel Mechanics of Adhesive Joints
,”
Polym. Eng. Sci.
,
17
(
7
), pp.
462
466
.
80.
Kaelble
,
D. H.
, and
Reylek
,
R. S.
,
1969
, “
Peel Adhesion—Rate Dependence of Micro Fracture Processes
,”
J. Adhes.
,
1
(
2
), pp.
124
135
.
81.
Chen
,
W. T.
, and
Flavin
,
T. F.
,
1972
, “
Mechanics of Film Adhesion—Elastic and Elastic-Plastic Behavior
,”
IBM J. Res. Dev.
,
16
(
3
), p.
203
.
82.
Kim
,
K.-S.
, and
Aravas
,
N.
,
1988
, “
Elasto-PIastic Analysis of the Peel Test for Thin Film Adhesion
,”
Int. J. Solids Struct.
,
24
(
4
), pp.
417
435
.
83.
Villey
,
R.
,
Creton
,
C.
,
Cortet
,
P.-P.
,
Dalbe
,
M.-J.
,
Jet
,
T.
,
Saintyves
,
B.
,
Santucci
,
S.
,
Vanel
,
L.
,
Yarusso
,
D. J.
, and
Ciccotti
,
M.
,
2015
, “
Rate-Dependent Elastic Hysteresis During the Peeling of Pressure Sensitive Adhesives
,”
Soft Matter
,
11
(
17
), pp.
3480
3491
.
84.
He
,
Y. F.
,
Wan
,
X. D.
,
Chen
,
Y. J.
, and
Yang
,
C. H.
,
2021
, “
Enhance the Debonding Resistance of Hydrogel by Large-Scale Bridging
,”
J. Mech. Phys. Solids
,
155
, pp.
1
16
.
85.
Niesiolowski
,
F.
, and
Aubrey
,
D. W.
,
1981
, “
Stress-Distribution During Peeling of Adhesive Tapes
,”
J. Adhes.
,
13
(
1
), pp.
87
98
.
86.
Karwoski
,
A. C.
,
2003
,
Testing and Analysis of the Peeling of Medical Adhesives From Human Skin
,
Virginia Tech
,
Blacksburg, VA
.
87.
Plaut
,
R. H.
,
2010
, “
Two-Dimensional Analysis of Peeling Adhesive Tape From Human Skin
,”
J. Adhes.
,
86
(
11
), pp.
1086
1110
.
88.
Zhang
,
L.
, and
Wang
,
J.
,
2009
, “
A Generalized Cohesive Zone Model of the Peel Test for Pressure-Sensitive Adhesives
,”
Int. J. Adhes. Adhes.
,
29
(
3
), pp.
217
224
.
89.
Poulard
,
C.
,
Restagno
,
F.
,
Weil
,
R.
, and
Leger
,
L.
,
2011
, “
Mechanical Tuning of Adhesion Through Micro-Patterning of Elastic Surfaces
,”
Soft Matter
,
7
(
6
), pp.
2543
2551
.
90.
Gent
,
A. N.
, and
Lindley
,
P. B.
,
1959
, “
Internal Rupture of Bonded Rubber Cylinders in Tension
,”
Proc. R. Soc. London Ser., A
,
249
(
1257
), pp.
195
205
.
91.
Shen
,
C. S.
,
Wang
,
H. F.
, and
Du
,
C. L.
,
2022
, “
Peeling of a Film From a Flexible Cantilever Substrate
,”
Mech. Res. Commun.
,
119
, pp.
1
10
.
92.
Roop
,
R. V.
,
Plaut
,
R. H.
,
Dillard
,
D. A.
, and
Ohanehi
,
D. C.
,
2002
, “
Peeling Pressure-Sensitive Adhesive Tape From Initially-Slack Thin Solid Film
,”
Proceedings of the 25th Annual Meeting of the Adhesion Society and the Second World Congress on Adhesion and Related Phenomena (WCARP-II)
,
Orlando, FL
,
Feb. 10–14
, pp.
129
131
.
93.
Steven-Fountain
,
A. J.
,
Atkins
,
A. G.
,
Jeronimidis
,
G.
,
Vincent
,
J. F. V.
,
Farrar
,
D. F.
, and
Chivers
,
R. A.
,
2002
, “
The Effect of Flexible Substrates on Pressure-Sensitive Adhesive Performance
,”
Int. J. Adhes. Adhes.
,
22
(
6
), pp.
423
430
.
94.
Plaut
,
R. H.
,
2010
, “
Peeling Pressure-Sensitive Adhesive Tape From Thin Elastic Strip
,”
J. Adhes.
,
86
(
7
), pp.
675
697
.
95.
Loukis
,
M. J.
, and
Aravas
,
N.
,
1991
, “
The Effects of Viscoelasticity in the Peeling of Polymeric Films
,”
J. Adhes.
,
35
(
1
), pp.
7
22
.
96.
Christensen
,
S. F.
, and
McKinley
,
G. H.
,
1998
, “
Rheological Modelling of the Peeling of Pressure-Sensitive Adhesives and Other Elastomers
,”
Int. J. Adhes. Adhes.
,
18
(
5
), pp.
333
343
.
97.
Afferrante
,
L.
,
Carbone
,
G.
,
Demelio
,
G.
, and
Pugno
,
N.
,
2013
, “
Adhesion of Elastic Thin Films: Double Peeling of Tapes Versus Axisymmetric Peeling of Membranes
,”
Tribol. Lett.
,
52
(
3
), pp.
439
447
.
98.
Zhu
,
Z. M.
,
Xia
,
Y.
,
Jiang
,
C. K.
,
Yang
,
Z. R.
, and
Jiang
,
H.
,
2021
, “
Investigation of Zero-Degree Peeling Behavior of Visco-Hyperelastic Highly Stretchable Adhesive Tape on Rigid Substrate
,”
Eng. Fract. Mech.
,
241
, pp.
1
12
.
99.
Ceglie
,
M.
,
Menga
,
N.
, and
Carbone
,
G.
,
2024
, “
Modelling the Non-Steady Peeling of Viscoelastic Tapes
,”
Int. J. Mech. Sci.
,
267
, pp.
1
10
.
100.
He
,
L.
,
Lou
,
J.
,
Chen
,
J.
,
Zhang
,
A.
, and
Yang
,
J.
,
2018
, “
An Adhesion Model for Plane-Strain Shearable Hyperelastic Beams
,”
Mech. Res. Commun.
,
90
(
1
), pp.
42
46
.
101.
Creton
,
C.
,
2003
, “
Pressure-Sensitive Adhesives: An Introductory Course
,”
MRS Bull.
,
28
(
6
), pp.
434
439
.
102.
Creton
,
C.
, and
Ciccotti
,
M.
,
2016
, “
Fracture and Adhesion of Soft Materials: A Review
,”
Rep. Prog. Phys.
,
79
(
4
), p.
046601
.
103.
Bekemeier
,
T.
,
Steve
,
C.
,
Dehlin
,
L.
,
Mitchell
,
T.
,
Ng
,
S.-S.
, and
Rich
,
D.
,
2020
, “Additive Organopolysiloxane Composition, Curable Composition, and Film,” U. S. Patent No. US20230192960A1.
104.
Dehlin
,
L.
,
Ferritto
,
M.
,
Mitchell
,
T.
,
Ng
,
S.-S.
,
Proffit
,
D.
,
Schmidt
,
R. G.
, et al
,
2022
, “Silicone Release Coating Composition and Methods for the Preparation and Use of Same,” U. S. Patent No. US20220380623A1.
105.
Li
,
Z.
, and
Zhenbin
,
N.
,
2020
, “Polyorganosiloxane Compositions Containing a 2-Substituted-1-Alkynyl-1-Cyclohexanol Useful as a Hydrosilylation Reaction Inhibitor,” U. S. Patent No. US10876019B2.
106.
Yamaguchi
,
K.
, and
Kishita
,
H.
,
2014
, “Diluent for a Flourine-Containing Silicone Coating Agent,” U. S. Patent No. US8765870B2.
107.
Newby
,
B. Z.
,
Chaudhury
,
M. K.
, and
Brown
,
H. R.
,
1995
, “
Macroscopic Evidence of the Effect of Interfacial Slippage on Adhesion
,”
Science
,
269
(
5229
), pp.
1407
1409
.
108.
Gordon
,
G. V.
, and
Schmidt
,
R. G.
,
2000
, “
PSA Release Force Profiles From Silicone Liners: Probing Viscoelastic Contributions From Release System Components
,”
J. Adhes.
,
72
(
2
), pp.
133
156
.
109.
Amouroux
,
N.
,
Petit
,
J.
, and
Leger
,
L.
,
2001
, “
Role of Interfacial Resistance to Shear Stress on Adhesive Peel Strength
,”
Langmuir
,
17
(
21
), pp.
6510
6517
.
110.
Leger
,
L.
, and
Amouroux
,
N.
,
2005
, “
Modulation of Adhesion at Silicone Elastomer–Acrylic Adhesive Interface
,”
J. Adhes.
,
81
(
10–11
), pp.
1075
1099
.
111.
Lamblet
,
M.
,
Verneuil
,
E.
,
Vilmin
,
T.
,
Buguin
,
A.
,
Silberzan
,
P.
, and
Leger
,
L.
,
2007
, “
Adhesion Enhancement Through Micropatterning at Polydimethylsiloxane-Acrylic Adhesive Interfaces
,”
Langmuir
,
23
(
13
), pp.
6966
6974
.
112.
Jones
,
D.
,
1997
,
Factors Affecting the Selection and Performance of Silicone Release Coatings
,
Dow Corning Corporation
,
Midland, MI
.
113.
ASTM-D6862
,
2021
,
Standard Test Method for 90 Degree Peel Resistance of Adhesives
,
American Society for Testing and Materials
,
West Conshohocken, PA
.
114.
D3330 A.
,
2018
,
Standard Test Method for Peel Adhesion of Pressure-Sensitive Tape
,
American Society for Testing and Materials
,
West Conshohocken, PA
.
115.
D6252 A.
,
2019
,
Standard Test Method for Peel Adhesion of Pressure-Sensitive Label Stocks at a 90 deg Angle
,
American Society for Testing and Materials
,
West Conshohocken, PA
.
116.
Zhenbin
,
N.
,
Fu
,
P.-F.
,
Mitchell
,
T.
, and
Wei
,
Y.
,
2023
, “Composition for Preparing a Release Coating, Release Coating Composition, and Related Methods,” U. S. Patent No. US20230050919A1.
117.
Stephen Edward
,
C.
,
David
,
R.
, and
Marc
,
T.
,
2007
, “Silicone Release Coating Compositions,” U. S. Patent No. US20070289495A1.
118.
Maiko
,
H.
,
Tsuyoshi
,
I.
, and
Mitsushi
,
Y.
,
2011
, “Protective Sheet and Use Thereof,” U. S. Patent No. US20110209994A1.
119.
Andriot
,
M.
,
Chao
,
S.
, and
Colas
,
A.
,
2009
, “Silicones in Industrial Applications: Silicones in Coatings,”
Silicon-Based Inorgic Polymers
,
Nova Science Publishers
,
Hauppauge, NY
, pp.
84
85
.
120.
Benedek
,
I.
,
2008
,
Manufacture of Pressure-Sensitive Products, Technology of Pressure-Sensitive Adhesives and Products
,
CRC Press
,
Boca Raton, FL
, pp.
10-1
10-65
.
121.
Owen
,
M. J.
, and
Dvornic
,
P. R.
,
2012
, “Surface Applications of Silicones,”
Surface Applications of Silicones
, Vol.
4
,
J.
Matisons
, ed.,
Springer Dordrecht
,
Dordrecht, Netherlands
, pp.
355
374
.
122.
Watson
,
A.
,
2020
,
Global Release Market Liner Annual Review
,
AWA Alexander Watson Associates B.V.
,
Amsterdam, Netherlands
, pp.
6
78
. https://awa-bv.com/market_report/awareness-report-industrial-release-liner-2020/
123.
Dudbridge
,
M.
,
2016
,
Handbook of Seal Integrity in the Food Industry
,
Wiley-Blackwell
,
Malden, MA
.
124.
Piergiovanni
,
L.
, and
Limbo
,
S.
,
2016
,
Food Packaging Materials
,
Springer
,
New York
.
125.
Hishinuma
,
K.
,
2009
,
Heat Sealing Technology and Engineering for Packaging: Principles and Applications
,
DEStech Publications, Inc
,
Lancaster, PA
.
126.
Halley
,
R. W.
, and
Malakoff
,
A. M.
,
2005
, “
A New High-Performance mVLDPE
,”
J. Plast. Film Sheet.
,
21
(
1
), pp.
13
26
.
127.
Gradys
,
A.
,
Sajkiewicz
,
P.
,
Minakov
,
A. A.
,
Adamovsky
,
S.
,
Schick
,
C.
,
Hashimoto
,
T.
, and
Saijo
,
K.
,
2005
, “
Crystallization of Polypropylene at Various Cooling Rates
,”
Mater. Sci. Eng.: A
,
413
(
414
), pp.
442
446
.
128.
Theller
,
H. W.
,
1989
, “
Heatsealability of Flexible Web Materials in Hot-Bar Sealing Applications
,”
J. Plast. Film Sheet.
,
5
(
1
), pp.
66
93
.
129.
Mazzola
,
N.
,
Cáceres
,
C. A.
,
França
,
M. P.
, and
Canevarolo
,
S. V.
,
2012
, “
Correlation Between Thermal Behavior of a Sealant and Heat Sealing of Polyolefin Films
,”
Polym. Test.
,
31
(
7
), pp.
870
875
.
130.
Morris
,
B. A.
,
2022
,
The Science and Technology of Flexible Packaging: Multilayer Films From Resin and Process to End Use
,
William Andrew
,
Norwich, NY
.
131.
ASTM, ASTM F88/F88M-23
,
2021
, “
Standard Test Method for Seal Strength of Flexible Barrier Materials
,” ASTM International, West Conshohocken, PA.
132.
Taheri
,
H.
,
Riggs
,
P.
,
Widem
,
N.
, and
Taheri
,
M.
,
2023
, “
Heat-Sealing Integrity Assessment Through Nondestructive Evaluation Techniques
,”
Packag. Technol. Sci.
,
36
(
2
), pp.
67
80
.
133.
Ilhan
,
I.
,
Turan
,
D.
,
Gibson
,
I.
, and
ten Klooster
,
R.
,
2021
, “
Understanding the Factors Affecting the Seal Integrity in Heat Sealed Flexible Food Packages: A Review
,”
Packag. Technol. Sci.
,
34
(
6
), pp.
321
337
.
134.
Konaganti
,
V. K.
,
Sadeghi
,
S.
, and
Goyal
,
S. K.
,
2018
, “
Thermo-Rheological Modeling and Simulation of Heat Sealing Process for Multi-Layer Flexible Packaging Applications
,”
Annual Technical Conference of the Society of Plastics Engineers
,
Orlando, FL
,
May 7–10
, pp.
2817
2824
.
135.
Market
,
M. A.
,
2021
, “
Medical Adhesives Market by Natural Type, Synthetic & Semi-Synthetic Resin Tpe (Acrylic, Silicone, Cyanoacrylate, pu, Epoxy), Technology (Water Based, Solvent Based, and Solids and Hot Melts), Application, and Region—Global Forecast to 2026
,” https://www.marketsandmarkets.com/Market-Reports/medical-adhesive-market-41880473.html, Accessed August 1, 2024.
136.
Ge
,
L.
, and
Chen
,
S.
,
2020
, “
Recent Advances in Tissue Adhesives for Clinical Medicine
,”
Polymers
,
12
(
4
), p.
939
.
137.
Richter
,
K.
,
Pinheiro
,
M.
,
Borcherding
,
K.
,
Hartwig
,
A.
,
Byern
,
J. V.
, and
Grunwald
,
I.
,
2019
, “Adhesives for Medical Applications,”
Green Chemistry for Surface Coatings, Inks and Adhesives: Sustainable Applications
,
R.
Höfer
,
A. S.
Matharu
, and
Z.
Zhang
, eds.,
The Royal Society of Chemistry
,
Cambridge, UK
, pp.
120
144
.
138.
Ebnesajjad
,
S.
,
2011
, “6—Adhesives for Medical and Dental Applications,”
Handbook of Polymer Applications in Medicine and Medical Devices
,
K.
Modjarrad
, and
S.
Ebnesajjad
, eds.,
William Andrew Publishing
,
Oxford
, pp.
103
129
.
139.
Venkatraman
,
S.
, and
Gale
,
R.
,
1998
, “
Skin Adhesives and Skin Adhesion: 1. Transdermal Drug Delivery Systems
,”
Biomaterials
,
19
(
13
), pp.
1119
1136
.
140.
Hwang
,
I.
,
Kim
,
H. N.
,
Seong
,
M.
,
Lee
,
S.-H.
,
Kang
,
M.
,
Yi
,
H.
,
Bae
,
W. G.
,
Kwak
,
M. K.
, and
Jeong
,
H. E.
,
2018
, “
Multifunctional Smart Skin Adhesive Patches for Advanced Health Care
,”
Adv. Healthc. Mater.
,
7
(
15
), p.
1800275
.
141.
Wong
,
S. H. D.
,
Deen
,
G. R.
,
Bates
,
J. S.
,
Maiti
,
C.
,
Lam
,
C. Y. K.
,
Pachauri
,
A.
,
AlAnsari
,
R.
,
Bělský
,
P.
,
Yoon
,
J.
, and
Dodda
,
J. M.
,
2023
, “
Smart Skin-Adhesive Patches: From Design to Biomedical Applications
,”
Adv. Funct. Mater.
,
33
(
14
), p.
2213560
.
142.
Czech
,
Z.
,
Kowalczyk
,
A.
, and
Swiderska
,
J.
,
2011
, “
Pressure-Sensitive Adhesives for Medical Applications
,”
Wide Spectra Quality Control
,
17
(
1
), pp.
310
332
.
143.
Rippon
,
M.
,
White
,
R.
, and
Davies
,
P.
,
2007
,
Skin Adhesives and Their Role in Wound Dressings
,
Wounds UK
,
London
,
3
, p.
76
.
144.
Singer
,
A. J.
,
Quinn
,
J. V.
, and
Hollander
,
J. E.
,
2008
, “
The Cyanoacrylate Topical Skin Adhesives
,”
Ame. J. Emergency Med.
,
26
(
4
), pp.
490
496
.
145.
McNichol
,
L.
,
Lund
,
C.
,
Rosen
,
T.
, and
Gray
,
M.
,
2013
, “
Medical Adhesives and Patient Safety: State of the Science: Consensus Statements for the Assessment, Prevention, and Treatment of Adhesive-Related Skin Injuries
,”
Orthop. Nurs.
,
32
(
5
), pp.
267
281
.
146.
Liu
,
J.
,
2012
, “Medical Adhesives and Skin Adhesion, Skin and Wound Care Division,” https://pstc.org/wp-content/uploads/2022/05/Liu-Jacob2012.pdf, Accessed 20, 2024.
147.
Kandur
,
M. Y.
,
Hemamalini
,
R.
, and
Öner
,
E. T.
,
2023
, “Global Industry Development and Analysis of Adhesives for Biomedical Applications,”
Adhesives in Biomedical Applications
,
Wiley
,
Hoboken, NJ
, pp.
25
46
.
148.
Group
,
I.
,
2023
, “
Bioadhesive Market: Global Industry Trends, Share, Size, Growth, Opportunity and Forecast 2023–2028
,” https://www.imarcgroup.com/bioadhesive-market, Accessed August 1, 2024.
149.
Benedek
,
I.
, and
Feldstein
,
M. M.
,
2009
,
Technology of Pressure-Sensitive Adhesives and Products
, 1st ed.,
CRC Press
,
Boca Raton, FL
, pp.
1
568
.
150.
Schott
,
H.
,
1971
, “
Contact Angles and Wettability of Human Skin
,”
J. Pharm. Sci.
,
60
(
12
), pp.
1893
1895
.
151.
Grove
,
G. L.
,
Zerweck
,
C. R.
,
Ekholm
,
B. P.
,
Smith
,
G. E.
, and
Koski
,
N. I.
,
2014
, “
Randomized Comparison of a Silicone Tape and a Paper Tape for Gentleness in Healthy Children
,”
J. Wound Ostomy Continence Nurs.
,
41
(
1
), pp.
40
48
.
152.
Mbithi
,
F.
, and
Worsley
,
P. R.
,
2023
, “
Adhesives for Medical Application-Peel Strength Testing and Evaluation of Biophysical Skin Response
,”
J. Mech. Behav. Biomed. Mater.
,
148
, pp.
1
8
.
153.
Hansen
,
D.
,
Zajforoushan Moghaddam
,
S.
,
Eiler
,
J.
,
Hansen
,
K.
, and
Thormann
,
E.
,
2020
, “
Performance of Polymeric Skin Adhesives During Perspiration
,”
ACS Appl. Polym. Mater.
,
2
(
4
), pp.
1535
1542
.
154.
Eiler
,
J.
,
Hansen
,
D.
,
Bingöl
,
B.
,
Hansen
,
K.
,
Heikenfeld
,
J.
, and
Thormann
,
E.
,
2020
, “
In Vitro Evaluation of Skin Adhesives During Perspiration
,”
Int. J. Adhes. Adhes.
,
99
, pp.
1
7
.
155.
Eiler
,
J.
,
Hansen
,
D.
,
Bingol
,
B.
,
Hansen
,
K.
, and
Thormann
,
E.
,
2020
, “
How the Viscoelastic and Sweat-Absorbing Properties of Skin Adhesives Affect Their Performance During Perspiration
,”
ACS Appl. Polym. Mater.
,
2
(
12
), pp.
5533
5541
.
156.
Lund
,
C. H.
,
Nonato
,
L. B.
,
Kuller
,
J. M.
,
Franck
,
L. S.
,
Cullander
,
C.
, and
Durand
,
D. K.
,
1997
, “
Disruption of Barrier Function in Neonatal Skin Associated With Adhesive Removal
,”
J. Pediatr.
,
131
(
3
), pp.
367
372
.
157.
Fumarola
,
S.
,
Allaway
,
R.
,
Callaghan
,
R.
,
Collier
,
M.
,
Downie
,
F.
,
Geraghty
,
J.
,
Kiernan
,
S.
, et al
,
2020
, “
Overlooked and Underestimated: Medical Adhesive-Related Skin Injuries
,”
J. Wound Care
,
29
(
Sup3c
), pp.
S1
S24
.
158.
Matsumura
,
H.
,
Imai
,
R.
,
Ahmatjan
,
N.
,
Ida
,
Y.
,
Gondo
,
M.
,
Shibata
,
D.
, and
Wanatabe
,
K.
,
2014
, “
Removal of Adhesive Wound Dressing and its Effects on the Stratum Corneum of the Skin: Comparison of Eight Different Adhesive Wound Dressings
,”
Int. Wound J.
,
11
(
1
), pp.
50
54
.
159.
Dugdale
,
D. S.
,
1960
, “
Yielding of Steel Sheets Containing Slits
,”
J. Mech. Phys. Solids
,
8
(
2
), pp.
100
104
.
160.
Xiang
,
Y.
,
Zhong
,
D.
,
Rudykh
,
S.
,
Zhou
,
H.
,
Qu
,
S.
, and
Yang
,
W.
,
2020
, “
A Review of Physically Based and Thermodynamically Based Constitutive Models for Soft Materials
,”
ASME J. Appl. Mech.
,
87
(
11
), p.
110801
.
161.
Mitsoulis
,
E.
,
2013
, “
50 Years of the K-BKZ Constitutive Relation for Polymers
,”
Int. Scholarly Res. Notices
,
2013
, pp.
1
22
.
162.
Bernstein
,
B.
,
Kearsley
,
E. A.
, and
Zapas
,
L. J.
,
1963
, “
A Study of Stress Relaxation With Finite Strain
,”
Trans. Soc. Rheol.
,
7
(
1
), pp.
391
410
.
163.
Rolón-Garrido
,
V. H.
, and
Wagner
,
M. H.
,
2009
, “
The Damping Function in Rheology
,”
Rheol. Acta
,
48
(
3
), pp.
245
284
.
164.
Wagner
,
M. H.
,
1979
, “
Elongational Behavior of Polymer Melts in Constant Elongation-Rate, Constant Tensile-Stress, and Constant Tensile Force Experiments
,”
Rheol. Acta
,
18
(
6
), pp.
681
692
.