Keywords

Introduction

As an typical laminated composite material, the standard PVB laminated glass comprises of a PVB interlayer sandwiched by two mono soda-lime and it has extensive applications in architecture, automobile industry, as well as the structural parts during the past few decades where they usually play an important role in human protection and structural integrity due to its excellent energy-absorbing and fragment-holding ability. However, mechanical behavior especially for the cracks of PVB laminated glass subject to loading has not been fully explored yet due to the complicity caused by the composite material, i.e., polymer material sandwiched by two pieces of brittle material. As a result, it becomes a pressing need to investigate the mechanical loading information contained in the macroscopic crack initiation and propagation which enables improved design or prevention for laminated PVB glass as a structural part.

In this chapter, there will be three methods mainly applied to investigate PVB laminated glass damage mechanism and crack growth: theoretical constitutive equations describing the failure criteria of PVB laminated glass under load, numerical simulations of crack initiation and growing on PVB laminated glass, experiments and tests of impact between impactor and laminated glass plate. For theoretical investigation, a damage-modified nonlinear viscoelastic constitutive equation is developed based on the updated Lagrangian approach. For numerical simulation investigation, a FEA model with proper material parameters verified by a classical example is established to describing the impact between pedestrian head and PVB laminated windshield. In addition, the extended finite element method (XFEM) is employed to study the low-speed impact-induced cracking in brittle plates in the present chapter. Upon head impact, the interaction between the stress field and the initiation and propagation of the radial and circumferential cracks are computed using XFEM. The effects of various impact conditions and sensitivity to initial flaw are discussed. The experiment analysis of PVB laminated glass is mainly based on quasi-static loading and dynamic impact. We carried out quasi-static Split Hopkinson Bar (SHPB) experiments to study the crack velocity and number of crack branching. For the dynamic out-of-plane impact tests carried out to study the crack propagation behavior of PVB laminated glass subjected to dynamic impact by using high-speed photography, the time histories of the averaged radial crack tip position, propagation velocity, and acceleration are recorded and the parametric study of impactor speed and mass, impact velocity, PVB thickness for PVB laminated glass are investigated. Then two different mechanisms of crack propagation on different glass layers are explained. Besides, a semiphysical model describing the relationship between the maximum cracking velocity and influential factors is established. In addition, the Weibull statistical model is suggested considering various factors to describe the macroscopic crack pattern thus providing a theoretical evidence for engineering practice. Finally, the relation between radial crack velocity and crack numbers on the backing glass layer and the relation between the crack length and the capability of energy absorption on the impacted glass layer are investigated.

Constitutive Relations

The constitutive relation of laminated composite material can well reflect the mechanical behavior of that. In order to study the crack in laminated composite material, the constitutive relation should be established firstly.

PVB windshield used in automotive industry are a kind of laminated composite material which comprise of a PVB interlayer sandwiched by two momo soda-lime glass sheet. Thus, pedestrian head impacts on the vehicle PVB windshield can be regarded as a scenario where a composite plate is subjected to a concentrated force by a spherical indenter on its top surface shown in Fig. 1. The composite plate consists of n = 2 individual brittle soda-lime glass layers of thickness tg and modulus Eg, sandwiched with a PVB interlayer of thickness tp and modulus Ep.

Fig. 1
figure 1

(a) PVB windshield crack morphology in real-world pedestrian-vehicle accident. (b) Schematic illustration of a sphere pedestrian headform impacting on a windshield glazing

By studying the constitutive relations of the PVB windshield, the crack in PVB windshield can be analyzed thoroughly. Lili et al. (Wang et al. 2003; Lili et al. 1991) study the mechanical behavior of polymer with performing dynamic impact and quasi-static tensile experiments. According to the experimental results, the behavior of the polymer can be described as the following constitutive relation:

$$ \sigma ={\sigma}_e+{\sigma}_{t_1}+{\sigma}_{t_2} $$
(1)

σt stand for the tensile stress.

$$ {\sigma}_e={\sigma}_m\left[1-\operatorname{exp}\left(-{\sum}_{t=1}^n\frac{{\left( m\varepsilon \right)}^t}{t}\right)\right] $$
(2)
$$ {\sigma}_{t_1}={E}_1{\int}_0^t\dot{\varepsilon}\left(\tau \right)\operatorname{exp}\left(-\frac{t-\tau }{\theta_1}\right) d\tau $$
(3)
$$ {\sigma}_{t_2}={E}_2{\int}_0^t\dot{\varepsilon}\left(\tau \right)\operatorname{exp}\left(-\frac{t-\tau }{\theta_2}\right) d\tau $$
(4)

where m, τ, E1,E2 are a nondimensional relative initial elastic modulus, relaxation time as defined in Maxwell body, elastic constants of first Maxwell body and elastic constants of second Maxwell body, respectively. θ1,θ2, are constants as well, standing for the relaxation times of first and second Maxwell body. Where σe stands for the stress with the damage material. σt1, σt2 standing for tensile stress of first and second Maxwell body. D is a “damage” variable.

Nevertheless, the above constitutive relations are only valid when strain is less than 7% (Lili et al. 1991), and it cannot express the relation completely. To further utilize the above relations, Fenghua et al. (Zhou et al. 1992) suggested the following equation to describe the constitutive relations:

$$ \sigma =\left(1-D\right){\sigma}_e+{\sigma}_{t_1}+{\sigma}_{t_2} $$
(5)

where D is a “damage” variable and is defined as:

$$ D=\left\{\begin{array}{cc}\hfill 0\hfill & \hfill \varepsilon \le {\varepsilon}_{th}\hfill \\ {}\hfill {\dot{D}}_0{\dot{\varepsilon}}^{\delta -1}\left(\varepsilon -{\varepsilon}_{th}\right)\hfill & \hfill \varepsilon >{\varepsilon}_{th}\hfill \end{array}\right. $$
(6)

And εth here regards as strain threshold of damage evolution. \( {\dot{D}}_0 \), δ both are material parameters determined by experiments. Then the dynamic damage criterion is put forward:

$$ {D}_f={\dot{D}}_0{\dot{\varepsilon}}^{\delta -1}\left(\varepsilon -{\varepsilon}_{th}\right) $$
(7)

Incremental Form of Constitutive Relations When the Material Is Not Damaged

When employing the constitutive relations, we should consider some conditions like whether the strain is larger than the threshold strain. Thus we need an equivalent strain to determine whether it is or not. As following is the equivalent:

$$ {\varepsilon}_{eq}=\frac{\sqrt{2}}{3}\sqrt{{\left({\varepsilon}_1-{\varepsilon}_2\right)}^2+{\left({\varepsilon}_2-{\varepsilon}_3\right)}^2+{\left({\varepsilon}_3-{\varepsilon}_1\right)}^2+\frac{3}{2}{\left({\gamma}_{12}^2+{\gamma}_{23}^2+{\gamma}_{31}^2\right)}^2} $$
(8)

where subscripts “1,” “2,” and “3” denote the directions of principal stresses.

On the basis of updated Lagrangian virtual work method, stress-strain relations of material that undergoes large deformation with the ignorance of physical nonlinear properties can be regarded as (Fischer and Washizu 1982):

$$ \left\{{S}_{ij}\right\}=\int_{0}^{t}E\left(t-\tau \right)\left[A\right]\frac{\partial \left\{{e}_{kl}\right\}}{\partial \tau } d\tau $$
(9)

{Sij}, {ekl} are Kirchhoff stress tensors and Green strain tensors, respectively.

$$ \begin{array}{l}\left[A\right]=\frac{1-v}{\left(1+v\right)\left(1-2v\right)}\cdot \\[6pt] {}\left[\begin{array}{cccccc}\hfill 1\hfill & \hfill \frac{v}{\left(1-v\right)}\hfill & \hfill \frac{v}{\left(1-v\right)}\hfill & \hfill 0\hfill & \hfill 0\hfill & \hfill 0\hfill \\ {}\hfill \frac{v}{\left(1-v\right)}\hfill & \hfill 1\hfill & \hfill \frac{v}{\left(1-v\right)}\hfill & \hfill 0\hfill & \hfill 0\hfill & \hfill 0\hfill \\ {}\hfill \frac{v}{\left(1-v\right)}\hfill & \hfill \frac{v}{\left(1-v\right)}\hfill & \hfill 1\hfill & \hfill 0\hfill & \hfill 0\hfill & \hfill 0\hfill \\ {}\hfill 0\hfill & \hfill 0\hfill & \hfill 0\hfill & \hfill \frac{1-2v}{2\left(1-v\right)}\hfill & \hfill 0\hfill & \hfill 0\hfill \\ {}\hfill 0\hfill & \hfill 0\hfill & \hfill 0\hfill & \hfill 0\hfill & \hfill \frac{1-2v}{2\left(1-v\right)}\hfill & \hfill 0\hfill \\ {}\hfill 0\hfill & \hfill 0\hfill & \hfill 0\hfill & \hfill 0\hfill & \hfill 0\hfill & \hfill \frac{1-2v}{2\left(1-v\right)}\hfill \end{array}\right]\end{array} $$
(10)

Similarly, as PVB windshield is a three-dimensional structure, Equations from 1 to 6 can be rewritten as (Swanson and Christensen 2015; Wang et al. 2007; Shen et al. 1987):

$$ {S}_{ij}={S}_{ij,e}+{S}_{ij,{t}_1}+{S}_{ij.{t}_2} $$
(11)

where

$$ {S}_{ij,e}={\sigma}_m\left\{1-\operatorname{exp}\left(-{mAE}_{kl}-{\sum}_{t=1}^n\frac{{\left({mE}_{ij}\right)}^t}{t}\right)\right\} $$
(12)
$$ {S}_{ij,{t}_1}={E}_1\int_{0}^{t}A\frac{\partial {E}_{kl}}{\partial \tau} \operatorname {exp}\left(-\frac{t-\tau }{\theta_1}\right) d\tau\qquad \quad\qquad\ \ \,$$
(13)
$$ {S}_{ij,{t}_2}={E}_2\int_{0}^{t}A\frac{\partial {E}_{kl}}{\partial \tau} \operatorname {exp}\left(-\frac{t-\tau }{\theta_2}\right) d\tau\qquad \quad\qquad\ \ \, $$
(14)

On the basis of Boltzmann superposition (Feng 1995), incremental form of constitutive relations can be obtained. Also, implementation in commercial FEA software needs the incremental form of the stress tensor of Eqs. 12, 13, and 14 to be rewritten. Thus, it is a must for us to reach the incremental form first. According to proofs of Chain Rule, Eq. 12 can be rewritten as follows:

$$ \begin{aligned}{dS}_{ij,e}&={\sigma}_m\operatorname{exp}\left(-{mAE}_{kl}-{\sum}_{t=2}^n\frac{{\left({mE}_{ij}\right)}^t}{t}\right) \\ {}&\quad\cdot\left[mA\frac{\partial {E}_{kl}}{\partial \tau }+{\sum}_{t=2}^n\frac{{\left({mE}_{ij}\right)}^{t-1}}{t}\frac{\partial {E}_{kl}}{\partial \tau}\right]\cdot \Delta t\end{aligned} $$
(15)

Considering Δt is short enough yields (Feng 1995):

$$ \frac{\partial {E}_{kl}}{\partial \tau }=\frac{\partial {E}_{kl}}{\partial t}=\frac{\partial {E}_{kl}}{\Delta \tau } $$
(16)

Thus Eq. 13 becomes:

$$ \begin{array}{l}\Delta {S}_{ij,e}={\sigma}_m\operatorname{exp}\left(-{mAE}_{kl}-{\sum}_{t=2}^n\frac{{\left({mE}_{ij}\right)}^t}{t}\right)\cdot \left[\!mA\Delta {E}_{kl}+{\sum}_{t=2}^n\frac{{\left({mE}_{ij}\right)}^{t-1}}{t}\Delta {E}_{kl}\!\right]\end{array} $$
(17)

Similarly, viscoelastic stress tensors for both low strain rate and high strain rate at t are:

$$ \Delta {S}_{ij,{t}_1}={E}_1{\theta}_1A\frac{1-\operatorname{exp}\left(-\frac{\Delta t}{\theta_1}\right)}{\Delta t}\Delta {E}_{kl} -\left(1-\operatorname{exp}{\left(-\frac{\Delta t}{\theta_1}\right)}^t\right){S}_{ij,{t}_1} $$
(18)
$$ \Delta {S}_{ij,{t}_2}={E}_2{\theta}_2A\frac{1-\operatorname{exp}\left(-\frac{\Delta t}{\theta_2}\right)}{\Delta t}\Delta {E}_{kl} -\left(1-\operatorname{exp}{\left(-\frac{\Delta t}{\theta_2}\right)}^t\right){S}_{ij,{t}_2} $$
(19)

Therefore, the incremental form of stress tensor during one unit time step Δt can be obtained:

$$ \begin{aligned}&\Delta {S}_{ij,n+1}=\Delta {S}_{ij,e}+\Delta {S}_{ij,{t}_1}+\Delta {S}_{ij,{t}_2}={\sigma}_m\!\operatorname{exp}\!\left(\!-{mAE}_{kl}-{\sum}_{t=2}^n\frac{{\left({mE}_{ij}\right)}^t}{t}\right) \\ & \cdot \left[mA\Delta {E}_{kl}+{\sum}_{t=2}^n\frac{{\left({mE}_{ij}\right)}^{t-1}}{t}\Delta {E}_{kl}\!\right]\!+{E}_1{\theta}_1A\frac{1-\operatorname{exp}\left(-\frac{\Delta t}{\theta_1}\right)}{\Delta t}\Delta {E}_{kl}\\ &-\left(1-\operatorname{exp}{\left(-\frac{\Delta t}{\theta_1}\right)}^t\right){S}_{ij,{t}_1}+{E}_2{\theta}_2A\frac{1-\operatorname{exp}\left(-\frac{\Delta t}{\theta_2}\right)}{\Delta t}\Delta {E}_{kl}\\ &-\left(1-\operatorname{exp}{\left(-\frac{\Delta t}{\theta_2}\right)}^t\right){S}_{ij,{t}_2}\end{aligned} $$
(20)

In the following way, the stress tensor can be calculated:

$$ {S}_{ij,n+1}={S}_{ij,n}+\Delta {S}_{ij,n+1}\quad \varepsilon <{\varepsilon}_{th} $$
(21)

where tn + 1 = tn + Δt.

Incremental Form of Constitutive Relations When the Material Is Damaged

When the material is not damaged, the incremental form of constitutive relations can be described using function relation above. However, when the strain is larger than the threshold strain, material turns into “damage” phase. As mentioned above, stress tensor in “damage” phase at time tn + 1 can be concluded as:

$$ {\left\{{S}_{ij}\right\}}_{n+1}={\left\{{S}_{ij}\right\}}_n+{\left\{\Delta {S}_{ij}\right\}}_{n+1} $$
(22)

and Eq. 21 can be rewritten as:

$$ {S}_{ij,n+1}={S}_{ij,n}+\Delta {S}_{ij,n+1} $$
(23)

Therefore,

$$ \begin{array}{l}{S}_{ij,n+1}={S}_{ij,n}+\Delta {S}_{ij,n+1}\\[5pt] {}=\frac{1}{1-{D}_{n+1}}{S}_{ij,n+1}^d+{1}^d-\frac{1}{1-{D}_n}{S}_{ij,n}^d\\[6pt] {}=\frac{1}{1-{D}_{n+1}}\Delta {S}_{ij,n+1}^d+\frac{\Delta {D}_{n+1}{S}_{ij,n}^d}{\left(1-{D}_{n+1}\right)\left(1-{D}_n\right)}\end{array} $$
(24)

where ΔDn + 1 = Dn + 1 − Dn is the increase of damage variable.

Rewriting Eq. 23 yields:

$$ \begin{aligned}\Delta {S}_{ij,n+1}^d &=\left(1-{D}_{n+1}\right)\Delta {S}_{ij,n+1}-\dfrac{\Delta {D}_{n+1}{S}_{ij,n}^d}{\left(1-{D}_n\right)}\\[8pt] {}&=\left(1-{D}_{n+1}\right)\Delta {S}_{ij,n+1}-\Delta {D}_{n+1}\Delta {S}_{ij,n}\end{aligned} $$
(25)

Similarly, the stress tensor under damage can be calculated as:

$$ {S}_{ij,\mathrm{n}+1}^d={S}_{ij,\mathrm{n}}^d+\Delta {S}_{ij,\mathrm{n}+1}^d\quad \varepsilon \ge {\varepsilon}_{th} $$
(26)

Verification of Constitutive Relations

The constitutive relation of PVB windshield are gotten preliminary and the accuracy of that should be studied next. In order to verify the above constitutive relations, a classical example is chosen. A thin plate with infinite length and 2b = 12 mm in width is under the uniform loadq(t) = 1 kN ⋅ m. There is a small crack in center with length 2a = 1 mm, see Fig. 2. Material properties are listed in Table 1.

Fig. 2
figure 2

Thin plate with center crack under uniform load

Table 1 Material parameters for plate

By implementing the above constitutive relations into FEA software numerical and adding the command of failure of material, results can be obtained. Comparing numerical results with theoretical ones, we can see that the constitutive relations results can coincide with the theoretical solutions well, shown in Fig. 3.

Fig. 3
figure 3

A comparison between theoretical solutions and FEA results

Numerical Simulation

PVB Windshield FEA Modeling Based on Constitutive Relations of PVB Laminated Glass

By verifying the constitutive relations above with a classical example, the relations have a sufficient accuracy. Then the FEA modeling of PVB windshield are introduced in detail following.

Contact Analysis

A nonconforming two-body contact model of a PVB windshield loaded by a headform (sphere) is illustrated in Fig. 4. The thickness of PVB windshield is 8.76 mm, the length of it is 350 mm, and the thickness of interlayer PVB is 0.76. Shell elements were chosen to mesh the two bodies in contact. Hourglass control and automatic surface- to-surface contact with dynamic friction coefficient 0.1 were employed in LS-Dyna. Initial nodal gaps between the two contact surfaces were prescribed. The contact elements were not activated until the penetration of the sphere into the composite glass occurred for the sake of saving computation time.

Fig. 4
figure 4

A simplified physical model of impact between head and windshield

An extra command that allows the failure of material is added in the input file of LS-Dyna. Then the above-mentioned constitutive relations with damage variable could be fully employed in FEA. The accuracy of the contact FEA model is firstly examined by classical Hertzian pressure calculation method. Illustration of classical Hertzian theory is shown in Fig. 5 (Daphalapurkar et al. 2007). The δ and a0 stand for mutual approach between two bodies and their contact radius.

Fig. 5
figure 5

A simplified crack-model based o n Hertzian pressure with moving coordinate at the crack tip is shown

Deviations between FEA and theoretical results are all within 1.5%. Comparison shown in Fig. 6 demonstrates the sufficient accuracy of the FEA model to stimulate the impact between pedestrian head and PVB windshield.

Fig. 6
figure 6

Comparison between FEA results and theoretical results

Material Model

The PVB laminated glass is m odeled according to the following principles: if the glass does not fail, the composite is treated as a shell; otherwise, the interlayer PVB film acts as a membrane. A Belytschko–Tsay shell element for the glass material and a membrane element for the interlayer are employed (Zhou et al. 1992; Fischer and Washizu 1982). Both types of elements are fully integrated in simulations so that no hourglass modes should be expected (Sharon and Fineberg 1999).

Glass and PVB film both can be treated as isotropic materials according to their physical and material properties. Therefore, we assume that each layer of composite windshield is isotropic with respect to its material symmetry lines and obeys Hooke’s law. It is assumed that, due to the influence of dynamic loading, there is no sliding between two contact layers. It then becomes reasonable to consider the composite windshield as integrity. As a result, an equivalent elastic modulus and Poisson’s ratio should be introduced as follows (Dwivedi and Espinosa 2003).

$$ \overline{E}=\frac{2{t}_g{E}_g+{t}_p{E}_p}{2{t}_g+{t}_p} $$
(27)
$$ \overline{v}=\frac{2{t}_g{v}_g+{t}_p{v}_p}{2{t}_g+{t}_p} $$
(28)

where \( \overline{E} \) and \( \overline{v} \) are equivalent Young’s modulus and equivalent Poisson’s ratio of windshield glazing.

Material parameters are listed in Table 2.

Table 2 Parameters used in both constitutive relations and finite element analysis

Results and Discussions

Internal Stress Analysis

The cracks in PVB windshield are demonstrations of internal and external stresses. Thus, the FEA model above is used to analyze the internal stresses. Only through investigating the internal stresses can we know the cracks on windshields better.

During the dynamic crack evolution shown in FEA analysis, stages of damage evolution could be sketchily drawn as Fig. 7. Three kinds of cracks are: circumferential cracks, web crack, and plastic crack. The circumferential cracks are mainly caused by pressure stress while the web cracks are mainly caused by tensor stress. Plastic cracks are the last stage that the windshield deforms plastically. In Fig. 8, three kinds of internal stresses, tensile stress σt, compressive stress σc, and shear stress σs, are plotted versus impact time. We can find out that the value of stresses sorted from large to small is σs, σc and σt. Lawn and Wilshaw (1975) stated that in addition to the Hertzian cone crack and the median vent crack, there is also an inelastic deformation zone where shear and hydrostatic compression is greatest. Quite similarly, it is obvious to see in Fig. 8 that shear stress is the greatest among three stresses, causing the windshield to deform plastically.

Fig. 7
figure 7

Diagram of different crack-growing stage

Fig. 8
figure 8

Three kinds of maximum internal stress in PVB windshield, tensile stress, compressive stress, and shear stress curves versus impact time

Internal Stress Variation Among Different Point on Windshield Plate

Internal stresses of different points on windshield plate are far different and crack information can prove it. The cracks become weaker as the radius grows larger for circumferential cracks, as do the radial cracks. In Fig. 9, maximum stresses appear at the very point that impacted by pedestrian head. Stresses decrease as the distances from the central point increase. The decreasing rates of three stresses are nearly the same according to the computational results.

Fig. 9
figure 9

Three kinds of maximum internal stress in PVB windshield, tensile stress, compressive stress, and shear stress curves versus the displacement from the central point

Effects of Poisson’s Ratio on Crack Angle

Poisson’s ratio plays an important role of cone crack angle. A parametric study is performed to manifest the effects on crack angle. Poisson’s ratio is ranged from 0.2 to 0.4, with a step-width of 0.05. Table 3 shows the results of FEA analysis compared to that of stress intensity factor analysis (SIF). Maximum difference between the results is below 7%. With the increase in Poisson’s ratio, the conical crack angle also increases though the increase pattern is still unknown.

Table 3 Comparison results of two models on Poisson’s ratio effect to crack angle
Effects of Impact Velocity

Vehicle speed is a critical factor in pedestrian-vehicle traffic accident. Extends and degrees of injuries largely depend on the impact velocity between pedestrian and vehicle. Therefore, studies of cracks on windshields under different head impact velocity would provide the essential foundations in pedestrian protection. According to the census data in NTADTU, in most cases, vehicles impact pedestrians at the speed of about 40–70 km/h. Thus, we vary the impact velocity of the head from 0 to 20 m/s and make impact velocity V0 as an independent variable. On the other hand, we choose the maximum radius of circumferential cracks Rc, the maximum length of radial cracks Lr, and the maximum radius of plastically deformed area Rp as dependent variables. Three dependent variables are illustrated and defined in Fig. 10.

Fig. 10
figure 10

Effect of initial impact velocity of the stress

In Fig. 11, the impact velocity has very great influence to three chosen variables. At the impact velocity of 20 m/s, Rc can reach up to about 642 mm, much larger than Rc ≈ 89 mm under the impact velocity of 5 m/s. Vehicle velocity plays an extremely important role in both pedestrian head injury and damage of windshield. In addition, it is reasonable to infer the extend of injuries of pedestrian head and impacting velocity of vehicle roughly through Rc, Lr, and Rp.

Fig. 11
figure 11

Effect of initial impact velocity to cracks

Numerical Simulation Based on the Extended Finite Element Method

After the head impacts on the windshield material, cracks including radial crack and circumferential crack appear in the glass material. The fracture characteristics of the windshield are widely recognized as one of the most important factors in automotive crashworthiness. The crack profiles (e.g., length, pattern, etc.) on the windshield material contain critical information for impact speed (which is extremely useful for accident reconstruction), vehicle crashworthiness, and insights for improving pedestrian and passenger protection. This motivates us to study the characteristics of low-speed head impact-induced crack propagation on the windshield glass material. In the past studies of impact on windshield glass, due to the intrinsic complexity of numerical analyses of cracking, researchers in automotive engineering often avoid explicit simulation of crack propagation. The following study aims to study the crack propagation characteristics (including both radial and circumferential cracks) when a model windshield undergoes low-speed impact and the numerical investigation of multiple crack propagation is based on the extended finite element method (XFEM) in this study.

Fundamentals of XFEM

XFEM incorporates a discontinuous displacement field across the crack facing away from the crack tip, in the form of:

$$ {u}^h=\sum\limits_{i=1}^n{N}_i(x)\left({u}_i+{b}_iH(x)+\sum\limits_{l=1}^4\sum\limits_{j=1}^2{c}_{il}{F}_l^i(x)\right) $$
(29)

where n is the number of nodes in the mesh, Ni(x) is the shape function of node i, ui are the classical DOFs of node i. bi and cil are the DOFs associated with the Heaviside “jump” function H(x), with value 1 above the crack and below the crack. The crack-tip function\( {F}_l^i(x) \) is:

$$ {F}_l\left(r,\theta \right)=\left(\sqrt{r} \operatorname {sin}\left(\frac{\theta }{2}\right),\sqrt{r} \operatorname {cos}\left(\frac{\theta }{2}\right),\sqrt{r} \operatorname {sin}\left(\frac{\theta }{2}\right)\operatorname{sin}\theta, \sqrt{r}\operatorname{cos}\left(\frac{\theta }{2}\right)\operatorname{sin}\theta \right) $$
(30)

where (r, θ) are the local polar coordinates at the crack tip.

Figure 12 illustrates the enriched nodes near a crack tip, where H(x) is discontinuous across the crack surface. Under general mixed mode-loadings, the asymptotic near-tip hoop and shear stress components are

Fig. 12
figure 12

Schematic of uniform mesh surrounding a crack used in XFEM simulation. Blue square nodes stand for the Heaviside function enrichment and red circular nodes stand for the crack tip enrichment. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article)

$$ \left\{\begin{array}{c}\hfill {\sigma}_{\theta \theta}=\frac{1}{4\sqrt{2\pi r}}\left({k}_1\left(3 \operatorname {cos}\left(\theta /2\right)\right.+\operatorname{cos}\left(3\theta /2\right)+{k}_{II}\left(-3 \operatorname {sin}\left(\theta /2\right)-3 \operatorname {sin}\left(3\theta /2\right)\right.\right)\hfill \\[7pt] {}\hfill {\sigma}_{\theta \theta}=\frac{1}{4\sqrt{2\pi r}}\left({k}_1\left(\operatorname{sin}\right.\left(\theta /2\right)+\operatorname{sin}\left(3\theta /2\right)+{k}_{II}\left(\operatorname{cos}\right.\left(\theta /2\right)-3 \operatorname {cos}\left(3\theta /2\right)\right)\hfill \\[7pt] {}\hfill {\sigma}_{\theta \theta}=\frac{1}{4\sqrt{2\pi r}}\left({k}_1\left(5 \operatorname {cos}\right.\left(\theta /2\right)-\operatorname{cos}\left(3\theta /2\right)+{k}_{II}\left(5 \operatorname {sin}\right.\left(\theta /2\right)-3 \operatorname {sin}\left(3\theta /2\right)\right)\hfill \end{array}\right. $$
(31)

For fracture propagation, we use the popular criterion of maximum local hoop stress, where the crack propagation direction is determined from Eq. 31 as

$$ {\theta}_c=arc \operatorname {cos}\left(\frac{3{k}_{II}^2+\sqrt{K_I^4+8{k}_I^2{k}_{II}^2}}{K_I^2+9{k}_{II}^2}\ \right) $$
(32)

Here, θc is measured with respect to a local polar coordinate system with its origin at the crack tip and aligned with the direction of the existing crack. Numerical simulations based on the above XFEM algorithm is carried out using commercial code ABAQUS.

Model and Methods

Model Setup

A featureless spherical headform is used to model the human head; The model head has a radius R = 90 mm and mass 4.5 kg. Note that effective head weights are changeable in different accident cases because human body may also get involved due to inertia, causing the effective impact mass larger than the mass of the pure head, especially during higher speed impact. Thus, the effective head mass M is also varied from 4.5 kg up to 90 kg by varying the density (from 1,500 to 30,000 kg/m3) and keeping volume the same in this study. According to 151 pedestrian–vehicle accident cases from the National Traffic Accident Database of Tsinghua University (NTADTU), most of the impact speeds of the vehicle in accident fall into the range of 40–80 km/h, i.e. 11.1–22.2 m/s, and that is the range of low-speed impact we consider. We focus on normal impact and assume that impact occurs at the center of the plate, we neglect the small windshield curvature and focus on a flat rectangular plate with dimensions a × b = 700 × 300 mm and thickness t = 4.76 mm as Fig. 13 shows.

Fig. 13
figure 13

Computational model setup. (a) Global view of the quarter model of head impact on a model windshield plate. (b) Dimensions of the featureless headform and windshield plate in a quarter model

In terms of the material properties, typical material parameters of glass are assigned to the plate, with density ρ = 2,500 kg/m3, Young’s modulus to the plate, Possion’s ratio v = 0.22, mode I energy release rate G1 = 10 J/m2, mode II and III energy release rate GII = GIII = 50 J/m2. The properties are assumed to be rate independent. To take advantage of symmetry, only one quarter of the model is needed for simulation. The boundary condition of the windshield is assumed to be fixed. The plate and head are modeled with eight-node linear brick elements. The contact between the head and plate obeys the Coulomb’s friction law with coefficient f = 0.1 (since according to our accident scene investigation, a small slip exists between head and windshield).

Computational Method

Suppose there is no crack or da mage, we first compute the deformation response with response to impact. For a given effective head mass M and impact speed v, the maximum displacement D is computed and presented in a dimensionless form in Fig. 14. For the range of impact speed and head mass considered in this study, D/t can be fitted as a dimensionless function of v/v0 and M/M0, as

$$ D/t=\left(135.3{\left(\nu /{\nu}_0\right)}^{0.7115}-{0.0305}^{518.5\left(\nu /{\nu}_0\right)}\right){\left(M/{M}_0\right)}^{0.2669} $$
(33)
Fig. 14
figure 14

Relationship between the initial impact velocity and displacement under different head masses; lines denote curve fitting, Eq. 33

where \( {v}_0=\sqrt{E/\rho } \) and M0 = abt ρ.

Since the impact speed is relatively low (much smaller than the wave speed), our simulation shows that the dynamic effect is relatively minor for the range of parameters investigated in this chapter. That is, for a given head mass, the computed stress field from dynamic impact (with incident speed v) is very close to that of quasi-static indentation (with maximum indentation depth D/t) as long as they induce the same deformation curvature of the plate (i.e., when v and D/t satisfy Eq. 33); see an example in Fig. 15 for the stress field. This enables us to simplify the dynamic impact problem into a quasi-static indentation one, where the normalized indentation depth D/t becomes the governing variable to indicate the effective “impact condition.”

Fig. 15
figure 15

Comparison between radial and hoop stress field resulted purely from impact deformation; to relieve the prominent stress, cracks need to form and radial crack is preferred to propagate before the circumferential crack owing to the more significant hoop stress field. (a) Contour plot of the radial stress field (unit: Pa). (b) Contour plot of the hoop stress field (unit: Pa)

Our previous study showed that for windshield fracture, circumferential cracks always grow after the radial crack has developed, except near the vicinity of impact where extensive damage can be found (Xu et al. 2009). This is consistent with Fig. 15 that the hoop stress field is more prominent than radial stress. Thus, the crack simulation can be decoupled into two steps. First, we embed a number of initial flaws at the center of the plate (impact location); this is consistent with the fact that extensive damage is initiated in the contact zone and that facilitates subsequent crack initiation and propagation. The size of the initial flaw is a0 = 3 mm with variable directions. When these initial flaws are very small, only the radial flaws will grow to become radial cracks upon impact, so as to release strain energy. For a given indentation depth D, the radial crack grows gradually until a certain length; with the increase of D, the final length of radial crack also increases. In the second step, based on the radially cracked specimen, new circumferential flaws with size a0 = 3 mm are introduced (the sensitivity of the flaw geometry and distribution is discussed below), and without further increasing the load, with the residual stress and strain fields inherited from the first step, the development of circumferential cracks is simulated by XFEM.

Rack Propagation Characteristics

Radial Crack Propagation Characteristics
Overall Crack Feature

In Fig. 16, for a typical displacement of D/t = 0.62, the evolved radial crack is shown along with the hoop stress field. Note that with the fully developed crack, this stress field is already much relaxed compared to its counterpart without cracking (see for example the difference between Figs. 15 and 16); moreover, the current hoop stress field is specified for the global coordinate, whereas during crack propagation, it is the local hoop stress that determines the crack trajectory. When the local maximum hoop stress exceeds a critical threshold, referred to as the damage stress σd, the crack grows incrementally toward such a local principal direction.

Fig. 16
figure 16

View of crack trajectory and the hoop stress field after the radial crack is fully developed at D/t = 0.62

In the beginning stage of impact, since the growth of radial crack can release more energy than the circumferential crack (due to the more prominent hoop stress field, Fig. 15), radial crack grows with the increase of the indentation depth D/t. It is noticeable that the crack growth direction keeps changing slightly with growth direction 22° approximately. This is mainly caused by the rectangular shape of the windshield plate, which makes the overall strain in the y-direction larger than that in the x-direction. When the crack grows further, the boundary constraint changes and the crack direction approaches about 45°.

During its growing process, the crack seeks to release as much energy as possible through mode I fracture. The released strain energy Estrain is computed as the difference between the strain energies of two specimens, one cracked and one without crack (at the same D/t). In Fig. 17, Estrain is normalized by the initial impact kinetic energy of the head Ekinetic (with M = 4.5 kg, the impact speed can be deduced from Eq. 33). It is also illustrated that the released strain energy is relatively small compared with the overall system energy, and it increases as D/t enlarges (although when it is normalized by the incident kinetic energy, which also increases as D/t gets larger, the dimensionless number Estrain/Ekinetic fluctuates somewhat).

Fig 17
figure 17

The normalized released strain energy during the propagation of radial and circumferential cracks. For radial crack, the released strain energy is a function of D/t (or equivalently the impact speed through Eq. 33); it is normalized by the incident kinetic energy with M = 4.5 kg. The circumferential crack propagates at D/t = 2.89 and it is normalized by the corresponding incident kinetic energy

Sensitivity of Initial Flaw and Fracture Criterion

The damage stressesσd of different soda-lime glass specimens are different due to its intrinsic stochastic flaws, usually from 10 to 60 MPa. Under quasi-static indentation, the effect of σd on crack evolution is shown in Fig. 18. The results indicate that as the critical stress for crack initiation σd becomes larger, the crack angle is smaller (the boundary constraint is more prominent) and the crack is shorter. In other words, if the impact condition is known, one can roughly estimate the damage stress threshold value according to the crack path and length. In what follows, σd = 10 MPa is employed.

Fig. 18
figure 18

Effect of different damage stresses on the radial crack trajectory, forD/t = 0.62

The model I energy release rate is another very important parameter governing the brittleness of the material. Consider typical material candidates for windshield such as soda-lime glass and PMMA, the former is a brittle material whereas the later has certain degree of ductility, their mode I energy release rates are about GI = 10 J/m2 and GI = 400 J/m2, respectively. Without loss of generality, we normalized the crack length C with plate size a. In Fig. 19, the effect on the resulting length of crack is examined by varying GI from 10 to 400 J/m2. Numerical study demonstrates that if the GI value is within the brittle material domain, there is little effect on the crack length; otherwise, if the windshield is made by the more ductile material, the crack length is shorter under the same plate deformation. In what follows, GI = 10 J/m2 is employed.

Fig. 19
figure 19

The effect of model I energy release rate effect on the radial crack length

The characteristics of initial flaw have some minor effects on crack propagation. We first keep the initial flaw length to be 3 mm and let the flaw angle θi to be 30°, 45°, and 60° with respect to the x-axis, and the final crack pattern is shown in Fig. 20. It is observed that θi does have some influence on the initial crack trajectory; however, eventually, all crack paths converge with similar lengths, indicating both the robustness of the current model. The length of the initial flaw has negligible effect on both the crack propagation pattern and crack length (as long as the initial flaw is short). When the initial flaw length is taken to be 0.5 mm, 1 mm, 3 mm, and 5 mm, the resulting radial crack lengths in these four scenarios are exactly the same; if the initial flaw is 10 mm, the crack propagation pattern is slightly different. Therefore, the 3 mm initial flaw used in the present study is validated and it is a relatively insensitive parameter in the present XFEM simulation.

Fig. 20
figure 20

The effect of the initial flaw angle hi on the radial crack pattern, forD/t = 0.62

Radial Crack Length

The length of the radial crack can be represented as a function of the quasi-indentation depth D/t, which is in turn related to the impact speed v for a given effective head mass (see Eq. 34). Figure 19 also shows the normalized crack lengths, i.e., C/a versus the normalized head displacement D/t. Initially, the crack length is almost proportional to head displacement; however, after a certain length owing to the fixed boundary, the crack length almost remains a constant. The relationship between C/a and D/t is nonlinear and may be fitted as

$$ \begin{array}{l}C/a=-0.4046+0.9231\left(D/t\right)-0.05074{\left(D/t\right)}^3\\ {}+0.03803\operatorname{exp}\left(D/t\right)-0.4243\sqrt{\left(D/t\right)}\operatorname{ln}\left(D/t\right)\end{array} $$
(34)

By combining Eqs. 33 and 34, a new method can be derived to estimate the impact speed directly based on the knowledge of radial crack length (assuming the effective head mass is known) using Newton–Raphson method, thus useful for traffic accident reconstruction. This is illustrated in the next section.

Equation 34 shows that if the head mass is fixed, the crack length increases nonlinearly with impact speed. In order to estimate the influence of the effective head mass, we can combine Figs. 14 and 19, along with Eqs. 33 and 34. For instance, from Fig. 14 (or Eq. 33), if the impact speed is fixed at 5 m/s, for different effective head masses, the variation of crack length (as a function of the corresponding D/t) can be obtained from either Fig. 19 or Eq. 34. Compared with the crack length of M = 22.5 kg, M = 45 kg, and M = 90 kg its 193.27%, 241.35%, and 294.23% for M = 4.5 kg, respectively. That is to say, the effective head mass has a strong influence over the crack length, indicating that during a low-speed impact accident, the shape of vehicle front end and pedestrian–vehicle contact point is critical in impact speed estimation since these factors will largely determine the pedestrian revolution angle. In addition, the influence of M becomes relatively smaller when M is large, implying that the influence of the effective mass starts to decay if the effective head mass is much larger than the windshield mass.

Circumferential Crack Propagation Characteristics

On the basis of the radially cracked geo metry, in the second step, we introduce circumferential flaws of length a0 = 3 mm along the radial crack, with a spacing of 50 mm (the spacing is found to be a relatively minor factor on the final circumferential crack pattern). The stress field calculated from the end of first step is transferred into the second step. Since the first step has relaxed most of the hoop stresses, the remaining excessive radial stress needs to be relaxed via the development of circumferential crack, and the most efficient way of releasing the radial stress is to develop the circumferential crack with maximum radius, i.e., on the same order of C. This is verified from Fig. 21: Cracks also grow near the contact zone because of the high stress. This is consistent with the observation that intensive cracks are identified within the contact area while the crack density is lower outside, and the largest circumferential crack almost bounds the radial cracks. The strain energy released by circumferential crack is given in Fig. 17, where the total released energy is less than 50% than that of radial crack. This evidence strongly supports the argument that crack is prone to grow in a radial way rather than a circumferential way, since more energy can be released. From Fig. 17, we can see that less than about 5% of the total kinetic impact energy is consumed by both radial cracks and circumferential cracks and thus there is enough room for improving the automotive windshield for enhanced energy mitigation efficiency and pedestrian protection (for example, using transparent energy absorption materials).

Fig. 21
figure 21

Final crack pattern (including both radial and circumferential cracks) on the model windshield plate subject to head impact with D/t = 0.62

Experiment Analysis

The mechanical behavior of PVB laminated glass is quite complex, especially under dynamic loadings due to the combination of rate-dependent effect of PVB and brittle-polymer composite structure. Therefore, experimental data are very important for the proper suggestion of constitution model and FE model to further investigation mechanical behavior of PVB laminated glass under both static and dynamic loadings.

Typically, one can categorize experiments for studying mechanical response of such a composite material into three regimes in terms of different strain rates: quasi-static, low impact speed, and high impact speed (Stout et al. 1999).

Experiment of Quasi-Static Loading

Because of the difficulties in c onducting a quasi-static tension experiment with a slippery surface and the windshield undergoes compression subject to human head impact, we adopted a compression experiment rather than the bending test which the shear effect will influence the mechanical property of the specimen to a large extent.

Experimental Setup

PVB laminated glass circular flake is introduced to be the testing specimen, with the dimension of u = 10 mm, where 0.76 mm-thick PVB interlayer is sandwiched by two pieces of 2 mm-thick glass, shown in Fig. 22. Note that the glass, as a typical rate-independent brittle material, has a density of ρ = 2,500 kg/m3, Young’s modulus of E = 70 GPa, and Poisson’s ratio of m = 0.22 (Forquin and Hild 2010). It is worthy to point out that PVB laminated glass is made by putting two glass sheets into an autoclave and compressed with 0.76 mm-thick of PVB interlayer under specific cycle of pressure (10 bar) and heat (120 °C). Thus, we can regard there is no slide between three layers (Wang et al. 2007). To make the experiment more engineering applicable, the thickness of specimen used here has the same thickness as that in most passenger cars. In experiment, specimen is put in a cylinder container which has a steel pin with a diameter u = 38 mm inside for compression and the experiments are carried out on Material Testing System (MTS). The load is applied perpendicularly through the pin to the sample. A clip-gage type extensometer is installed between the gaps to measure the displacement of the pin as a function of load during the experiment, shown in Fig. 23. 1 mm/min, 0.1 mm/min, and 0.01 mm/min are three loading rates, corresponding to the strain rate on the upper and lower sides of the specimen of \( \dot{\varepsilon}\approx 1\times {10}^{-3}{s}^{-1} \), \( \dot{\varepsilon}\approx 1\times {10}^{-4}{s}^{-1} \), \( \dot{\varepsilon}\approx 1\times {10}^{-5}{s}^{-1} \) chosen to measure the quasi-static mechanical property of PVB laminated glass. During tests, unloading threshold value is set to be 80% of the maximum loading value and thus complete load–displacement history is recorded. The temperature and relative humidity during the experiments remain in the range of 25–26 °C and 40–45%, respectively.

Fig. 22
figure 22

Illustrations and dimensions of specimen used in both quasi-static and dynamic impact experiments. (a) Illustration of the testing specimen before test. (b) Detail dimension of specimen

Fig. 23
figure 23

Quasi-static compression experimental setup on MTS. (a) Compression experiment setups. (b) Principles of compression

Experiment Results

It is commonly accepted that the failure process of composite subject to quasi-static compression loading usually involves a sequential damage accumulation process (Xu et al. 2009). As aforementioned, we set the unloading threshold value to be 80% of the maximum load value due to the major load decrease occurred during the experiment when major crack could be observed clearly. Thus, we define the stress and strain at this moment as major failure on-set (MFO) stress and strain. Figure 24 shows us the “load decrease” phenomenon and the morphology of the cracked samples.

Fig. 24
figure 24

Major load decrease occurrence with the onset of major crack

Figure 25 illustrates three different stress–strain relation curves under three loading rates. The stress–strain curves show “nonlinearity” characteristic in mechanical behavior of PVB laminated glass and this “nonlinearity” phenomenon is probably caused by both nonlinear mechanical response of PVB interlayer and progressive microcrack growth in sample. As the loading rate increasing, the MFO strain increases while the MFO stress remains nearly the same. The major responsible reasons are: (i) in extremely low strain rate (quasi-static) situation, the outer glass panel plays a critical role in mechanical response and (ii) glass is a rate-independent material whereas PVB is a rate-dependent one.

Fig. 25
figure 25

Stress–strain curves obtained from quasi-static compression experiments under three different loading rates, i.e., 1 mm/min, 0.1 mm/min, and 0.01 mm/min

Number of Crack Branching

The numbers of cracks in testing samples according to different quasi-static loading rates are illustrated in Fig. 26. The number of complete cracks (or major cracked branching) increases with the loading rate. As aforementioned, in quasi-static situation, the stress remains the same whereas the strain rises with the increasing loading rate. Therefore, the number of cracked branching, namely the “damage degree,” depends on the value of strain, not the stress, in quasi-static circumstance that is in accordance with Wang’s (Wang et al. 2006) study. Further, if a nonlinear fitting is tried on the number of crack branching versus loading rate, we will harvest the following explicit formula with fitting coefficient R = 0.999.

$$ N=\left[8.975{W}^{0.339}\right] $$
(35)
Fig. 26
figure 26

Number of cracked branching at different loading rates

where [X] is Round function, e.g. [5.2] = 5. N is the number of crack branching and W is loading rate with unit of mm/min. From Eq. 35, we can infer the numbers of crack zones under different loading rates in the quasi-static domain.

Mandelbrot (1977) pointed out that crack branching contained critical information of crack propagation speed and complexity. Bouchaud et al. (2012) further investigated the relationship between crack length and crack branching mode and concluded

$$ {D}_f=\frac{\operatorname{lg}\;b}{\operatorname{lg}\;k} $$
(36)

where \( {D}_f=\frac{\operatorname{lg}\;b}{\operatorname{lg}\;k} \) is degree of branching, b is the number of basic branching pattern, and K ≈ 0.4 in brittle material.

Assuming that the initial crack branching length is a0 and the length becomes a1 = a0 + ka0 (Bouchaud et al. 2012), thus after the initial crack branches for i times, the length finally comes to

$$ {a}_i=\left(1+\cdots +{K}^i\right){a}_0 $$
(37)

Since the number of branching can be written as N ¼ biþ1, we are able to know that

$$ i={\mathrm{log}}_b^N-1 $$
(38)

According to Freund’s (Michel and Freund 1990) theory , the dynamic crack propagation velocity in a finite body obeys the following relation:

$$ v={C}_r\left(1-\frac{a_0}{a_i}\right) $$
(39)

where Cr is the Rayleigh wave velocity and v is the crack propagation velocity.

By combination of Eqs. 36, 37, 38, and 39, the relation between crack velocity and crack branching number is

$$ v={C}_r\frac{1+{b}^{-1/{D}_f}-2{N}^{-1/{D}_f}}{2\left(1-{N}^{-1/{D}_f}\right)} $$
(40)

Further, we substitute Eq. 35 into Eq. 40 and get

$$ v={C}_r\frac{1+{b}^{-1/{D}_f}-2{\left[8.975{W}^{0.339}\right]}^{-1/{D}_f}}{2\left(1-{\left[8.975{W}^{0.339}\right]}^{-1/{D}_f}\right)} $$
(41)

Equation 40 may help us to clarify the crack propagation speeds in different loading rates. We may also conclude that crack propagation speeds are different during the three loading rates in quasi-static experiments based on Eq. 41.

Experiments of Dynamic Out-of-Plane Loading

By starting from a fundamental point, quasi-static mechanical behavior of PVB laminated glass has been widely investigated experimentally above (Xu et al. 2011a). Unlike quasi-static cracking problem, dynamic fracture is more complex and challenging but there are few reported investigations that focused on the dynamic fracture of PVB laminated glass except for some numerical simulations those based on probabilistic damage mathematical model (Forquin and Hild 2010), continuum damage mechanics (Zhao et al. 2005, 2006a, b; Sun et al. 2005), explicit finite element method (Bois et al. 2003; Timmel et al. 2007; Sun et al. 2009), and extended finite element method (Xu et al. 2010) that often neglected the important PVB layer and most of them did not involve explicit crack growth and pattern analysis. The dynamic experimental investigation is pressing need.

System Setup of Dynamic Out-of-Plane Loading

The impact system setup is sketched in Fig. 27, where a weight block can slide freely along the two standing poles. At the end of the drop-weight, there is a tip which is used to hit the force direction converter on the testing sample and it can be referred in Fig. 28. It is a necessity to convert the vertical impact direction into a horizontal one, which would facilitate the capture of the film.

Fig. 27
figure 27

(a, b) Schematic of the drop-weight tower experiment setup. The testing sample is illustrated along with the force direction convertor

Fig. 28
figure 28

The schematic illustration of the entire kit of experiment setup. (a) A schematic illustration of the basic layout of high-speed photography experiment setup. (b) Detail illustration of optical setup

Upon impact high speed photography is employed for capturing the dynamic crack growth in the plate in situ. The entire system consists of a high-voltage charging controller, multispark box, two field lenses, and an array camera with films (see in Fig. 28). In addition, a multispark high-speed camera is adopted to provide enough spot light source. The multispark box provides 16 independently spot lights triggered by a high-voltage charging controller by a preset adjustable interval time from 1 to 9,999 μs whose charging voltage is able to reach up to 30 kV. Meanwhile, a 4 × 4 array camera with 16 films is set correspondingly on the other side to receive the light and therefore record the crack image information. In order to make sure that the images recorded on the films are in the time series of the ignition of the spot lights, two field lenses are used to make up the optical path (see Fig. 28b). Moreover, the delay controller is introduced to control the time interval between each spot light. In addition, to minute the exact ignition time for each spot light, the electric-light receiver along with the oscillograph monitor are used together where the electric-light receiver is employed to collect the electric-light signals and send them to the oscillograph monitor. Reference control experiments of impact-induced fracture are carried out using pure glass sheets, with results in good agreement with the literature (Nielsen et al. 2009; Sharon and Fineberg 1999) (see in Fig. 29), thus validating the current test system.

Fig. 29
figure 29

Comparison of the crack velocity propagation speed of pure glass between experiments results extracted from the current one and the one in Nielsen et al. (2009). Note that the external impact energy is 1 J in the current experiment while it is 0.2 J in the experiment in Nielsen et al. (2009)

The Preliminarily Experiment to Investigate the Radial Crack Propagation

Experiment Condition

In the preliminarily impact fracture experiment to investigate the radial crack propagation behavior of PVB laminated glass , the minimum mass of the drop-weight is 0.5 kg and the maximum height of the drop-weight is 1.84 m (corresponds to the maximum impact velocity of 6 m/s); both the mass and height of the drop-weight are also varied in this experimental investigation. Prior to impact, the drop-weight is suspended and fixed through electromagnet. The temperature and relative humidity during the experiments remain in the range of 20–21 and 20–25% respectively.

Specimen Preparation

The standard PVB layer has a Young’s modulus EPVB = 100 MPa, Poisson’s ratio υPVB = 0.48, and mass density ρPVB = 870 kg/m3 (Xu and Li 2009). To make close connection to the automotive industry, the PVB thickness is tPVB = 0.76 mm for passenger cars. In addition to the middle PVB layer, for the upper and lower soda-lime glass sheets in laminated glass, the thickness tglass = 2 mm, with ρglass = 2,500 kg/m3, Eglass = 70 GPa, and υglass = 0.22. They are bonded together at 10 bar and 120 °C following standard procedures, and the effective in-plane dimension is 200 × 150 mm.

Experimental Results and Discussion
Crack Morphology

A series of experiments under different loading conditions, i.e., with different height and mass of the drop-weight are recorded. Figure 30 shows the typical dynamic crack propagation snapshots of the PVB laminated glass sheet, in response to a 1 kg weight dropped from 1 m height. Extensive radial fracture is observed, with no appearance of circumferential crack, which is consistent with numerical simulation that when a foreign object impacts a large windshield, the radial crack always appears before the circumferential crack (Xu et al. 2010). It is seen that there are about ten radial cracks on the composite sheet, and the development and evolution of each crack are quite different due to the inevitable stochastic flaw in glass. Here, the average crack length is employed to discuss the behavior of radial crack propagation. By using nonlinear fitting method to the experimental points obtained from the experiments, we finally got the crack tip position in time history. Then the averaged time evolution history of the radial crack tip velocity and acceleration can be obtained from the time history of the radial position of the crack tip as showed in Fig. 31.

Fig. 30
figure 30

Snapshots of crack propagation with 1 kg of drop weight and 1 m of drop height

Fig. 31
figure 31

Averaged crack tip velocity and acceleration in time history for the case in Fig. 28

The measured crack velocity first increases and then fluctuates slightly, which goes accordance with the experiment results of cracking speeds of pure soda-lime glass at various cracking modes in Nielsen et al. (2009) and Sharon and Fineberg (1999), as well as being consistent with previous numerical simulation results of several brittle materials (Song et al. 2008; Grégoire et al. 2009) and composites (Daphalapurkar et al. 2007; Dwivedi and Espinosa 2003).

As proposed by Ravichandar and Knauss (1984), near the vicinity of crack tip, there is a microcrack zone which impedes the further propagation of crack. Within the impact area, the spacing between the initial cracks is so small and such an intensively damaged zone may reduce the cracking speed substantially. With the advancement of cracks, the retardment effect will be greatly lessened until it reaches an almost constant influence, when the crack speed becomes steadier in the later stage. Thus, the crack evolution may be qualitatively divided into two stages: in stage I, the cracking speed keeps rising, and in stage II the crack speed becomes more stable but with some fluctuations. The transition occurs roughly at about time (further discussion will be made later). During stage I, the crack propagation acceleration is decreasing until stage II begins; that is, as the stress in the crack tip accumulates and develops, finally the crack tip will be able to break through the energy barrier surrounding material to reach the equilibrium.

In this example, the averaged steady-state cracking speed during stage II is about \( {\overline{v}}_{PVB}\approx 811\ m/s \). Comparing with the stable cracking speed of Polymethyl Methacrylate (PMMA)\( {\overline{v}}_{PMMA}\approx 600\;m/s \) (Sharon and Fineberg 1999) and tempered soda-lime glass \( {\overline{v}}_{GLASS}\approx 1470\;m/s \) (Nielsen et al. 2009) (One may refer to Fig. 32 for detail comparison. Note that although the impact energy is a little bit different in (Xu and Li 2009) and this experiment, the comparison is still valid), the PVB laminated glass may be regarded a composite material whose overall effective “fracture toughness” is between that of PMMA and pure glass. In Fig. 32, we can also see that the general variation trend of crack speeds in two different materials are almost the same. One may use the Rayleigh wave speed vR of the glass sheet to normalize the crack speed (since cracking occurs in the glass sheet), i.e., vR = 3,098 m/s. Thus, the stable cracking speed of PVB laminated glass is about 0.26 vR in this particular case, which is lower than the typically measured maximum speed of about 0.4–0.5 vR in brittle materials and we can clearly see that the PVB interlayer serves to slow down the crack speed, which would also make the material as well as the brittle-polymer sandwiched structure a candidate for energy absorption/shielding.

Fig. 32
figure 32

Comparison of the crack velocity propagation speed among pure glass, PVB laminated glass, and PMMA. Note that data for PMMA is extracted from

Different Driving Mechanisms of In-Plane Cracking on Two Glass Layer

Figure 33 shows a typical crack propagation comparison on supported and loaded glass layers at the loading speed with PV B thickness h = 0.76 m. Here, time t = 0 μs is used to denote crack tip initiation for the radial cracks on supported glass layer. It is clear that the radial cracks (Fig. 33a–d) initiated firstly on the “supported glass layer,” then the radial cracks on the “loaded glass layer” generated after about t = 600 μs as shown in Fig. 33f. On the other hand, it is also interesting to observe that the cracks on the two glass layers are completely overlapped even though they propagate at totally different times.

Fig. 33
figure 33

Selected sequence of images depicting the radial crack growth on both glass layers

In fact, the stress wave dominated inertial effects during impact loading on the PVB laminated glass play an important role in crack propagation (Singh and Parameswaran 2003; Clements et al. 1996). Due to the high modulus of glass material and the fixed mode of the specimen, the plate is small enough to ignore the bending deformation during the impact. For the convenience of analysis, a schematic diagram in a layer system consisting of two glass plates bonded by a PVB interlayer is shown in Fig. 34. Impact occurs on the surface of the “loaded glass layer” of the PVB laminates, the compressive waves caused by the impact travel down the depth direction (Fig. 34). Due to its viscoelasticity, the PVB layer plays an important role in reducing the amplitude and the speed of waves that eventually pass through by internal friction (Thom 2005), compared with the glass layer. Finally, the compressive waves reflect from the boundaries as tensile waves, which continuously arrive at the supported plate thus the tensile dominated cracks initiate while reaching the limit of dynamic fracture toughness KID of the material (Lambros and Rosakis 1997). Considering that glass plate is more sustainable in compressive stress wave but prone to fail in tensile stress wave, the radial cracks initiate first on supported glass layer.

Fig. 34
figure 34

Schematic of the impact process in a layer system consisting of two glass plates bonded by a PVB interlayer. The cracks propagate into the loaded glass layer by reinitiation from a surface flaw

Effect of Drop Weight and Height

Next, we investigate the influence of drop weight and height on the crack propagation. The drop mass ranges from 0.5 to 2 kg, and the height varies from 0.6 to 1.4 m. Table 1 shows the detail combination of the parametric experiments. Figures 7 and 8 show that with the increase of drop weight and height, the basic trend of crack would remain almost the same while the value would increase, i.e., larger mass and higher impact speed will cause faster crack propagation. Zhang et al. (2010) investigated the effect of loading rate on crack velocities in concrete and found with the increase of loading rate, the crack velocity increase proportionally; in quasi-static indentation, it was also found that with larger indentation load, the radial and lateral cracks in ceramic also became longer. All these conclusions are qualitatively consistent with our experimental findings. Particularly, \( {\overline{v}}_{PVB} \) increases 17.70% and 12.21% when the impact speed changes from 3.43 to 4.43 m/s and 5.24 m/s, respectively, and \( {\overline{v}}_{PVB} \) increases 26.72% and 14.92% when the drop weight is increased from 0.5 to 1 kg and 2 kg, respectively. Furthermore, if we combine the effect of impact mass and velocity together into the term of impact energy E, and normalize it by a reference potential energy En = 9.81 J, then the normalized steady-state cracking speed, i.e., \( {\overline{v}}_{PVB}/{v}_R \) can be fitted as

$$ {\overline{v}}_{PVB}/{v}_R=0.167\pm 0.137/\left(1\pm 10 \operatorname {exp}\left(1.50\right.\left(0.739-E/{E}_0\right)\right) $$
(42)

for the range of parameters in this study. From Fig. 35, it can also be found that the stable crack velocity will not likely to increase indefinitely as the external impact energy gets higher, and it has a theoretical limit, i.e., Rayleigh wave speed vR, but not reaching it.

Fig. 35
figure 35

The relationship between the steady-state crack propagation velocity and the external impact energy

Regarding the transition time between stages I and II, the general trend is that the smaller the external energy, the faster the transition time (although the difference is relatively small and still on the same order of magnitude). That is to say, if the impact energy is higher, it is easier for the crack speed to achieve steady-state. The detailed physical mechanism of the variation of such a transition time will be explored in future.

The Further Experiment Investigating the Radial and Circular Crack

Experiment Condition

In the further experiment to investigate the radial and circular crack propagation of PVB laminated glass, a weight block with mass of 2 kg is first evaluated to a certain height and then released to create impact loading. The height of the drop-weight adopted in our work varies from 300 to 900 mm. An impactor is fixed at the tip of the force direction converter to control the contact shape and area. Different shapes may be designed for the impactor, and in this study, the impactor shape is confined to be cylindrical and the top is hemispherical. The diameter of the cylindrical is 10 mm and the total length of the impactor is 25 mm. The mass of the impactor (including the force direction converter) is 0.545 kg. In addition, a force sensor with sufficient accuracy and frequency is attached right below the drop-weight to record force–time history and the signals recorded by the force sensor are sent to the oscillograph monitor to record force–time history.

Specimen Preparation

The plate is consisted of a PVB interlayer sandwiched by two brittle glass sheets (For PVB interlayer: Young’s modulus Ep = 0.1 GPa, short time shear modulus G0 = 0.33 GPa, long time shear modulus G1 = 0.69 MPa, bulk modulus K = 20 GPa, Poison’s ratio vp = 0.49, density ρp = 1,100 kg/m3, and decay factor β = 12.6 s−1; For glass: Young’s modulus, Poison’s ratio and thickness of the glass. Eg = 70 GPa, Poison’s ratio vg = 0.22 and thickness of the glass is 2 mm. The thickness of the PVB interlayer varies from 0.38 to 3.04 mm during our parametric experimental study (where thickness of 0.76 mm is for common passenger cars). The specimen is prepared using the same manufacturing process (with compression at 10 bar and 120 °C) as that used in automotive windshield. Glass is a typical linear elastic brittle material where its facture strain is about 0.1% while the PVB is a rubber-like material which may sustain a much larger deformation during tension. The unique design enables the possible protection against impact with small amount of energy and dissipates larger impact energy through the large deformation of interlayer.

The specimen is clamped within two metal cover sheets with thin layers of rubber pad inside to avoid possible scratches and stress concentration on the sample surface and distribute the boundary force more uniformly. The tightening torque for the screw bolt is set to be 4 Nm which properly mimics the clamping boundary condition. Six screws penetrating the metal cover sheets are fixed to make sure the metal cover sheets are placed in right horizontal position and reduce the rebound under impact. A force direction conversion part (indicated in Fig. 27) is employed to transfer the vertical impact force into a horizontal one, providing an out-of-plane dynamic loading to the sample.

Experimental Results and Discussion
Crack Morphology

Two kinds of macroscopic cra ck patterns, radial crack and circular crack, are observed and all the pieces of glass plates after impact are stay connected to the PVB layer. Hoop stress is larger than radial stress such that radial crack appears before circular crack, therefore, it gives us an opportunity to capture and study the evolution history of radial and circular cracks separately. Figure 36 is typical crack propagation process which contains a series of selected sequence of images depicting the radial and circular crack growth in PVB laminated glass specimen respectively. Time t = 0 μs is used to denote crack tip initiation for both radial and circular cracks. The crack propagation velocity v is determined as the crack length increase ∆l over the time increase ∆t. For radial crack, at time ti, multiple major crack tips conform approximately a circle with radius ri on average, shown in Fig. 36a. Therefore, radial crack propagation velocity vr = Δl/Δt = (ri + 1 − ri)/(ti + 1 − ti). For circular crack, the total length of n main cracks at ti is added and divided by the crack number as \( {\overline{l}}_i \) and such that circular crack propagation velocity \( {v}_c=\varDelta l/\varDelta t=\left({\overline{l}}_{i+1}-\overline{l_i}\right)/\left({t}_{i+1}-{t}_i\right) \). In addition, the B-spline interpolation method is used to smooth the experimental points in this study.

Fig. 36
figure 36

Selected sequence of images depicting the crack growth at the impact velocity of 3.13 m/s, with PVB thickness of 0.76 mm. (a) Radial crack. (b) Circular crack

Crack Propagation

To enhance the credibility of the experi mental results, the dynamic strain indicator is introduced. Locations near the force sensor and the top of the impactor are fitted with strain gages respectively so as to monitor in situ stains (and stresses) transmit during the impact process. Thus, we could estimate the delay of the impact force from the force sensor to the glass surface and depict the accurate force–time curve acting directly on the glass surface (see dashed line in Fig. 37). Figure 37 shows a typical crack propagation velocity time history curves for radial and circular crack propagation velocities respectively in response to 2 kg weight from 500 mm drop height (impact velocity: 3.13 m/s) to a specimen with 0.76 mm PVB interlayer. Time t = 0 μs is employed to denote crack tip initiation for radial crack. Therefore, the negative time correspond to preinitiation conditions (see Fig. 37). Note that the crack only exists in the glass sheets such that the stress component normal to the crack front σ11 is the major driving force for crack initiation and propagation. The dynamic crack propagation velocity v may be expressed as

$$ {v}^2=\frac{\sigma_{11}+{\sum}_{11}}{\rho_g\left(1+{\upvarepsilon}_{11}\right)} $$
(43)
Fig. 37
figure 37

Load–time curve and crack growth velocity history of the radial crack and the circular crack, in response to 500 mm drop height (impact velocity:3.13 m/s) and 0.76 mm PVB thickness

where the contact stress at the crack tip which can be calculated based on existing dynamic contact model by attaching a strain gage near the impactor. ∑11 is the strain component. By supposing the proportionality coefficient D between σ11 within the framework of linear elasticity (Zhang et al. 2010), the above equation may be rewritten as

$$ {v}^2=\frac{1+D}{\rho_g\left(1/{\sigma}_{11}+1/E{}_g\right)} $$
(44)

In this impact loading condition, the maximum crack propagation velocity is 990 m/s and 790 m/s for radial and circular cracks, respectively. As Xu et al. (2010) indicated that σ11 is larger in hoop stress than circumferential stress so radial crack propagates faster than circular crack. The radial crack propagation velocity is lower than that in pure soda-lime glass (Nielsen et al. 2009) where no obvious circular crack occurs. The PVB interlayer acts as an energy-absorbing material thus to reduce the crack propagation velocity. Meanwhile, PVB interlayer provides a connection media for cracked glass panel such that circular crack could continue to dissipate impact energy after the full growth of radial cracking.

Impact occurs on the outer surface of the glass plate at around −250 μs (see Fig. 37). A compressive loading wave front travels down at around 5,300 m/s (Zhao et al. 2006b) in the glass layer. The PVB layer presents the dual advantages of delaying the passage of the elastic wave into the backing glass layer and reducing the amplitude and the wavelength of wave that eventually passes through by internal friction caused by its viscoelasticity. Both of these effects are beneficial in utilizing the energy absorbing capabilities of the PVB interlayer to the fullest. The compressive waves arrive at the free surface of the backing layer and reflect as tensile waves. Thus, the delay of the crack initiation probably results from the transmitting and reflection of the stress wave and the increasing process of the stress intensity factor caused by increasing elastic energy stored near the crack tip. Similar phenomena have also been observed in Lambros and Rosakis (1997). More and more tensile waves continually arrive at the backing plate and thus the radial cracks initiate while reaching the limit of dynamic crack initiation toughness \( {K}_{IC}^d \). For the rate-sensitive material, the PVB interlayer is rate dependent while the fracture occurs in the glass plate which is rate independent. Considering the fact that cracks grow in the glass layer, therefore, the impact velocity has little effect on the \( {K}_{IC}^d \) value, which means the effect of rate sensitivity can be ignored. The propagation of the radial crack can be divided into three stages (see Fig. 37). After initiation, the crack tip accelerates from about 400 m/s to around 990 m/s within 30 μs. We can see that the impact force remain increasing after the initiation of the radial crack, which means the elastic energy stored near the crack tip keeps increasing thus causing the acceleration sequences. Then the crack velocity remains constant at about 990 m/s during the next 25 μs. However, it decelerates to less than 600 m/s in under 100 μs after initiation. The entry of crack tip into the region of compressive stress as it approaches the boundaries and the arrival of less stress wave at the crack tip are two responsible reasons for the deceleration phenomenon (Lambros and Rosakis 1997). Through further investigation, the reason for the deceleration and crack arrestment phenomenon is relevant to the boundaries. The history of the radial crack velocity at various impact velocities from 2.42 to 4.2 m/s with the PVB thickness of 0.76 mm is plotted as a function of crack length in Fig. 38 and it is clearly shown that the crack velocity start decreasing at the same position.

Fig. 38
figure 38

Velocity–length curve of the radial crack at various impact velocities from 2.42 to 4.2 m/s with the PVB thickness of 0.76 mm

For the circular crack, the crack initiation occurs 1,750 μs after the impact on the glass surface (see Fig. 37). Combined with the images depicting the radial and circular crack growth, it is believed that circular cracks initiate long after the radial crack growth as we stated before. An observation of the specimen, at the end of the test, showed that circular cracks merely existed on the loaded layer, according to which we can conclude that the circular cracks nuclear under the influence of surface acoustic waves (namely “Rayleigh waves”). The propagation of this kind of elastic waves is limited to the loaded surface of the solids. The Rayleigh waves spread at a much lower speed compared with the longitudinal compression waves. Besides, the lateral dimensions of the sample are much larger than those in longitudinal direction. Therefore, the circular crack appears generally later than the radial crack which is confirmed by the previous finding (Xu et al. 2010). The circular crack propagates under the tensile stress caused by the reflected tensile Rayleigh waves, which results in that the circular cracks always emerge first close to the boundaries. The propagation of the circular crack can be divided into three stages as well (see Fig. 37). Both the first and the third stage contain the acceleration, the maintenance, and the deceleration process similar as those in radial cracks. There is more space for circular crack to grow than radial crack such that the total crack lengths are longer in circular cracks. Besides, the circular crack has the lower peak value. However, between the two stages, there exists a special stage which could be called “slowing-down stage.” In this stage, the crack propagation velocity decelerates to almost 0. This is a direct consequence of the fracture at the glass interface caused by the radial crack. We can clearly see that the circular crack propagation always delays at the relatively large fracture where the stress waves have to propagate through the interlayer, without which the circular crack cannot transfer effectively across the interface (Park and Chen 2011).

Effect of Impact Velocity

To investigate the effect of impact velocity vd, the both radial and circular crack propagation characteristics, respectively, series of parametric experiments are conducted. Figure 7 shows the crack propagation velocity time history of both radial and circular cracks under various impact velocities from 2.42 to 4.2 m/s (drop height from 300 to 900 mm). Two kilogram drop weight and 0.76 mm PVB thickness are employed in this study. Three repeated experiments under each condition are carried out to ensure the results accuracy (see error bars in Fig. 39). Time t = 0 μs is employed to denote crack tip initiation. For the radial crack, in the case with higher impact velocity, a shorter acceleration time is needed to reach the top propagation velocity where the top value is also higher. This is consistent with the fact that at higher impact velocity with larger wave amplitude, the load applied to the specimen increases at a faster rate and thus crack acceleration can be larger during the dynamic loading process. Besides, additional experiments also indicate that the higher the impact velocity, the shorter delay time is needed before initiation (see Fig. 40). However, the case with higher impact velocity would also decelerate faster and earlier due to the limited available cracking space in the panel plane. Similar to the radial crack, higher impact velocity results in the shorter initiation time, the shorter acceleration time, and the larger peak value for the circular crack. To establish the model combining the crack velocity with the impact conditions, we chose the variable “the maximum velocity” to quantify our study, which apparently depends on both the impact conditions and the material property. We can suppose that the maximum velocity vmax at the crack tip during the whole propagation process has an implicit relationship with the impact force, the duration of the loading time and the dynamic initiation fracture toughness of the material in the present investigation. According to the principle of dimensionless unifying, the equation of vmax could have the following form as:

$$ {v}_{\mathrm{max}}=\frac{{AP}_d^{2/3}}{{\left({k}_{IC}^d\right)}^{2/3}{t}_d} $$
(45)
Fig. 39
figure 39

Crack velocity history of the radial and circular crack at various impact velocities from 2.42 to 4.2 m/s with the PVB thickness of 0.76 mm

Fig. 40
figure 40

Initiation time of the radial and circular crack at various impact velocities from 3.13 to 4.2 m/s with the PVB thickness of 0.76 mm

where pd is the average contact force over the contact period and td is the duration of the loading, respectively, A is a macroscopic crack pattern dependent coefficient. According to the momentum theorem, we have:

$$ {p}_d{t}_d={m}_d{v}_d $$
(46)

This means that the maximum velocity at the crack tip can be expressed in terms of the impact velocity vd and the drop weight md

$$ {v}_{\mathrm{max}}=\frac{{Am}_d^{2/3}{v}_d^{2/3}}{{\left({k}_{IC}^d\right)}^{2/3}{t}_d^{5/3}} $$
(47)

Thus, the maximum velocity contains two model parameters: the dynamic initiation fracture toughness \( {k}_{IC}^d \), which can be geometry dependent, and the macroscopic crack pattern dependent parameter A > 0. Both parameters are subject to the experimental data via best-fitting.

Effect of PVB Thickness

Series of parametric experiments are conducted to study the effect of PVB thickness tp on the crack propagation characteristics as well. Three different PVB thicknesses, i.e., 0.38 mm, 0.76 mm, and 1.52 mm are chosen for the parametric experiments. Two kilogram drop weight is employed. The crack propagation velocity time history of both radial and circular cracks under three different PVB thicknesses can be seen in Fig. 41. The impact velocity used to investigate the radial crack is 2.42 m/s (i.e., drop height is set to be 300 mm) while the impact velocity is 3.13 m/s (i.e., drop height is set to be 500 mm) for circular crack. Martinez et al. (1998) reported that the capability for transmitting the impact energy depends on the thickness and type of the adhesive used. Therefore, the variation in PVB thickness will have a great effect on delaying the passage of the elastic wave into the backing glass layer and reducing the amplitude and the wavelength of wave through internal friction (Xu et al. 2011b), which results in reducing the maximum radial crack speed as show in Fig. 41. Considering the wave speed in the through thickness direction of around 5,320 m/s in the glass layer and 300 m/s in the PVB interlayer (which means the PVB interlayer plays an important role in slowing down the stress wave), the transit time is around 3 μs, 6 μs, and 11 μs, respectively with the PVB thickness ranged from 0.76, 1.52, to 3.04 mm. Compared with the delay time in the order of 102 μs for radial crack initiation in Fig. 10, we can conclude that the delay of the radial crack initiation mainly results from the increasing process of the stress-intensity factor. The thicker PVB thickness would reduce the amplitude of elastic wave crossing the PVB interlayer by internal friction thus to weaken the increasing process of the stress-intensity factor. Therefore, in the case with thicker PVB layer, a longer delay time is needed to initiate. For the circular crack caused by Rayleigh wave, similar phenomenon is observed. Considering the Rayleigh surface wave speed in the in-plane direction of around 3,370 m/s in the glass layer (Sharon et al. 2002), the transit time is around 37 μs, which is much longer than the transmit of the elastic wave in the medium. Similarly, the delay of the circular crack initiation mainly also results from the increasing process of the stress-intensity factor. In addition, from the force–time curve, we can see that the compressive waves stay around the loaded side for a relatively long time during the impact, which also have an effect on the increasing process of the stress-intensity factor.

Fig. 41
figure 41

Crack velocity history of the radial and circular crack with various PVB thicknesses from 0.38 to 1.52 mm at the impact velocity of 2.42 m/s and 3.13 m/s, respectively

It is clearly seen that the glass sheet with thicker PVB layer starts at a higher initiation velocity v0 for both crack patterns. Since the exact expression of v0 is not clarified yet, we may implicitly represent the expression to the initiation velocity v0 as v0(tp). Due to the fact that \( {k}_{IC}^d\propto {\sigma}_c\propto {v}_0 \) (see Eq. 44), where σc is the fracture stress, we have the following expression for the initiation velocity (Berezovski and Maugin 2007).

$$ {v}_0^2\left({t}_p\right)={c}_R^2\left(1+D\right)\left(1-{\left(1+\frac{{\left({k}_{IC}^d\right)}^2}{M}\right)}^{-1}\right) $$
(48)

where M is a constant coefficient and CR is the Rayleigh wave velocity. In this case, we arrive at an expression for the dynamic initiation fracture toughness in Eq. 47 in the form of

$$ {k}_{IC}^d=\sqrt{\frac{{Mv}_0^2\left({t}_p\right)}{C_R^2\left(1+D\right)-{v}_0^2\left({t}_p\right)}} $$
(49)

Therefore, we may come to the expression for the maximum velocity at the crack tip as follows:

$$ {v}_{\mathrm{max}}=\left[\frac{{Bm}_d^2{v}_d^2{t}_d^{-5}\left({c}_R^2\left(1+D\right)-{v}_0^2\left({t}_p\right)\right)}{v_0^2\left({t}_p\right)}\right] $$
(50)

where B > 0 is another constant coefficient depending on the macroscopic crack pattern. According to Eq. 50, the thicker the PVB layer, the lower of the maximum velocity for both radial and circular crack, which is consistent with the experiment result (see Fig. 41). This is expected since the glass sheet with thicker PVB layer would absorb more impact energy via molecule vibration and friction by the interlayer as polymer-like material, verified by Xu et al. (2011c). As the Rayleigh wave velocity CR ≈ 3370 m/s (Hauch and Marder 2010), which is sufficiently high for the initiation velocity vo. The expression may be simplified as

$$ {v}_{\mathrm{max}}={\left[\frac{{Bm}_d^2{v}_d^2{t}_d^{-5}\left({c}_R^2\right.\left(1+D\right)}{v_0^2\left({t}_p\right)}\right]}^{1/3}=c{\left[\frac{m_d^2{v}_d^2{t}_d^{-5}}{v_0^2\left({t}_p\right)}\right]}^{1/3} $$
(51)

Then the agreement between theory and experiment may be achieved by an appropriate choice of the values of the constant coefficient c (c1 and c2 for radial and circular crack, respectively). The fitted models for the maximum velocity of radial crack \( {v}_{\mathrm{max}}^r \) and circular crack \( {v}_{\mathrm{max}}^c \) are in the following form of:

$$ \left\{\begin{array}{l}{v}_{\mathrm{max}}^r\approx {c}_1{\left[\frac{m_d^2{v}_d^2{t}_d^{-5}}{v_0^2\left({t}_p\right)}\right]}^{1/3}=0.0657{\left[\frac{m_d^2{v}_d^2{t}_d^{-5}}{v_0^2\left({t}_p\right)}\right]}^{1/3}\\ {}{v}_{\mathrm{max}}^c\approx {c}_2{\left[\frac{m_d^2{v}_d^2{t}_d^{-5}}{v_0^2\left({t}_p\right)}\right]}^{1/3}=0.0545{\left[\frac{m_d^2{v}_d^2{t}_d^{-5}}{v_0^2\left({t}_p\right)}\right]}^{1/3}\end{array}\right. $$
(52)

The fitted data and experimental data are shown together in Fig. 42 (the experimental data in the figure is the average maximum crack velocity). The good agreement between the two proves the model to be dependable.

Fig. 42
figure 42

Experiment data and fitted data of the maximum crack velocity from the semiphysical with various impact velocities from 2.42 to 4.2 m/s with PVB thickness from 0.38 to 1.52 mm. (a) Radial crack. (b) Circular crack

Statistical Analysis on Macroscopic Cracking Morphology

To confirm the crack prop agation mechanism analysis before and understand the crack propagation mechanism from another perspective, the relationship between the number of cracks and the impact velocity or the PVB thickness is established. Since all cracks locate in the surface of glass sheet and glass is a material with internal stochastic flaws which may serve as crack nuclei, a statistical model may be needed to describe the two macroscopic crack patterns. Table 4 shows the detail combination of the statistical experiments. One hundred repeated experimental data for each specified condition have been used for each experiment condition to obtain a statistical result. The two-parameter Weibull model is employed to study the distribution of both radial and circular crack number under each experiment condition. Weibull statistics are commonly used in the engineering community, especially in the fracture statistics. In the specific case of Weibull statistics analysis in this study, the cumulative probability function is written as follows:

$$ {p}_f=1-\operatorname{exp}\left(-{\left(\frac{N}{N{}_0}\right)}^m\right) $$
(53)
Table 4 Parametric statistical experiments: linear regression estimates of the Weibull parameters under different experiment conditions

where Pf is the cumulative probability, N is the number of radial or circular crack (Nr or Nc), m and N0 is the distribution shape parameter and characteristic crack number, respectively. To improve the statistical accuracy, the number of circular crack is counted with decimal during the statistical process (that is the number of circular crack with incomplete circle is determined by the length ratio of incomplete circle and the complete one). For the ease of accessing information, the double logarithm of the Weibull Eq. 55 is used in number analysis, yielding the Weibull parameters in a simple graphical representation of the data. Figure 43 shows the graphs with X = 1 nN plotted on the horizontal axis and Y = ln[ln(1/(1 − pf))] on the vertical axis. One can perform a simple linear regression analysis to get the Weibull parameters under different experiment conditions. The slope of the line is the Weibull modulus, m. The characteristic crack number, N0 is the value of crack number for which ln[ln(1/(1 − pf))] is zero. In this study, the linear regression analysis is used for the Weibull parameter estimates under each experiment condition (see Table 4).

$$ \operatorname{ln}\left(1-{p}_f\right)=-{\left(\frac{N}{N_0}\right)}^m $$
(54)
$$ \operatorname{ln}\left[\operatorname{ln}\left(\frac{1}{1-{p}_f}\right)\right]=m \operatorname {ln}\;N-m \operatorname {ln}\ {N}_0 $$
(55)
Fig. 43
figure 43

Fitted Weibull model plots for macroscopic cracking morphology: (a) radial crack at various impact velocities from 2.42 to 4.2 m/s; (b) circular crack at various impact velocities 2.42–4.2 m/s; (c) radial crack with different PVB thicknesses from 0.38 to 3.04 mm; and (d) circular crack with different PVB thicknesses from 0.38 to 3.04 mm

Firstly, the effect of different impact velocities on the distribution of both radial and circular crack number is investigated. Figure 44a shows the probability density curves of radial and circular crack number under different impact velocities. The characteristic crack number indicates the distribution location along the X axis and is expected to vary according to the impact velocity or the PVB thickness. The distributions for both radial and circular crack number are proved to move to the right (higher value) as impact velocity increases. This is reasonable because higher impact velocity produces more elastic energy, and thus it is necessary to create larger new surface to release the excess energy. In addition, the probability density curves considering different PVB thicknesses to describe the radial and circular crack number are shown in Fig. 44b. Contrary to the effect of impact velocity, the distributions for both radial and circular crack number move to the left (lower value) as PVB thickness increases since thicker PVB layer absorbs more elastic energy as stated before. This statistical method could also be applied to the engineering applications and play an important role in the laminated glass design.

Fig. 44
figure 44

Probability density curves of the radial and circular crack number under various boundary conditions. (a) For the radial and circular crack number at four different impact velocities, i.e., 2.42 m/s, 3.13 m/s, 3.7 m/s, and 4.2 m/s. (b) For the radial and circular crack number with four different PVB thicknesses, i.e., 0.38 mm, 0.76 mm, 1.52 mm, and 3.04 mm

The Relation Between Radial Crack Velocity and Crack Numbers on the Backing Glass Layer

Here, the effect of the radial crack number on their own crack propagation velocity as well as the crack generation on the other glass layer is thoroughly investigated. Firstly, the radial crack velocity on the backing glass layer with different crack numbers is calculated. A theoretical model from the perspective of energy conversion is established to depict the cracking process of backing glass layer and elucidate the fundamental reason that causes the variation of the crack velocity.

The specimen in response to impact speed V = 3.7 m/s is chosen for 100 repeated experiments by considering the intrinsic stochastic flaws in glass laminated samples. A series of images depicting the crack growth on both glass plates are recorded at the set time intervals. Typical crack propagation process is shown in Fig. 45. The glass plate directly contacting with the impactor is defined as “impacted glass layer,” while the glass plate on the other side is referred as “backing glass layer.” t = 0 μs represents the time when impact is triggered. Figure 45ad demonstrate the whole cracking process on the backing glass layer where cracks always initiate first, and only radial crack pattern is observed while both radial and circular cracks appear on the impacted glass layer (under the effect of Rayleigh waves, whose propagation is limited to the loaded surface of the solids) long after the cracking of the backing one (Fig. 45ef). However, the radial crack number on the backing glass plate varies from 18 to 112 within 100 repeated experiments, which is quite a large deviation. Further, it is also qualitatively observed that the number of radial cracks have great effect on the later crack generation on the impacted glass layer as well as the propagation characteristics of radial crack on the backing glass layer itself. In the following section, the effect of the different radial crack numbers of backing glass layer would be thoroughly discussed.

Fig. 45
figure 45

Images depicting the crack growth on the two glass layers (V = 3.7 m/s)

The crack velocity history on the backing glass layer with different radial crack number is calculated separately. All the radial cracks present approximately the same length during the propagation, conforming a circle overall (Fig. 45). Therefore, the crack velocity vr is calculated as the radius increase ∆r over the time increase ∆t, i.e., Δrt. Figure 46 shows the crack velocity variation with the crack length lr under three selected radial crack numbers (Nr = 18, 64, and 108). Crack propagations far from the boundaries and the center of the glass plate are focused to eliminate the possible boundary effect (i.e., the crack length between 20 and 40 mm). Results indicate that the number of radial cracks on the backing glass layer regularly negatively influences cracking propagation velocity themselves, i.e., the crack velocity vr always decreases with the radial crack number increasing.

Fig. 46
figure 46

Radial crack velocity versus crack length with different radial crack number (V = 3.7 m/s)

During the propagation of the cracks on the backing glass layer, the impacted glass layer remains intact as stated before (Fig. 46). However, due to the adhesion by PVB interlayer, the two glass layers have the same radius of deformed region rf (Fig. 46). The bending energy of the impacted glass plate caused by the impact can be estimated by assumption that the backing glass layer is still intact, in which case the two glass plates can be regarded as “single one with 2hg thickness.” Therefore, the bending energy of impacted glass layer is equal to half the energy in the single glass plate with 2hg thickness. After the contact between the impactor and the glass plate, kinetic and bending energies balance ρgΩV2 ∼ Eg(2hgκ)2Ω (Vandenberghe et al. 2013), where \( \Omega =2{h}_g\pi {r}_f^2 \) is the volume of deformed region and V is the impact speed whose variation can be ignored during such a short period. Thus, the curvature \( \kappa \sim {w}_0/{r}_f^2\sim V/2{h}_g{c}_g \), yielding rf ∼ (2hgcgt)1/2, where w0 = Vt is the indentation and \( {c}_g=\sqrt{E_g/{\rho}_g} \) is the sound speed in the glass material. The bending energy of impacted glass plate Ub1 can be expressed as

$$ {U}_{b1}={k}_1\pi {E}_g\left(\frac{V}{c_g}\right){r}_f^2{h}_g $$
(56)

where k1 is a coefficient.

Considering the low Young’s modulus of PVB material (i.e., Ep = 0.1 GPa), the bending energy of PVB Ub3 is about 0.001 of that in the glass plate, which has been proved by the finite element simulation with a three-dimensional laminated plate FE model using 1 × 1 mm surface size element. The boundary condition of the glass plate model is set to be clamped, which is consistent with the real experimental condition. Thus, Ub3 is small enough to be ignored compared to that of Ub1.

While for the backing glass layer in the cracking process with crack length lr, the number of radial cracks Nr on the backing layer has been set and would not change as the cracks extend. Thus, the bending energy of the backing glass plate Ub2 can be estimated by the energy of Nr triangular beams of length lr with the neglect of transverse bending (Vandenberghe et al. 2013):

$$ \begin{array}{l}{U}_{b2}={E}_g{h}_g^3{w}_0^2{N}_r\operatorname{tan}\left(\frac{\pi }{N_r}\right)/3{l}_r^2\\ {}\approx {k}_2\pi {E}_g{h}_g{r}_f^2{\left(\frac{V}{c_g}\right)}^2\left(\frac{1}{2}+\frac{\pi^2}{6{N}_r^2}\right)\end{array} $$
(57)

where k2 is another coefficient. Therefore, the bending energy of the whole laminated plates Ub can be estimated as

$$ \begin{array}{l}{U}_b=\left({U}_{b1}+{U}_{b2}\right)\\ {}\approx \pi {E}_g{h}_g{r}_f^2{\left(\frac{V}{c_g}\right)}^2\left[{k}_1+{k}_2\left(\frac{1}{2}+\frac{\pi^2}{6{N}_r^2}\right)\right]\end{array} $$
(58)

From the perspective of energy transferring, during the cracking process, the elastic energy of the laminated plates (i.e., the bending energy) caused by the impact is converted into the glass surface energy through the fracture behavior. Here we define the increasing rate of the plates elastic energy as “ηb,” and the energy release rate is “ηc.” Therefore, we have

$$ {\eta}_b=\frac{dU_b}{dt}\approx \pi {E}_g{h}_g^2{\frac{V}{c_g}}^2\left[{k}_1^{\prime }+{k}_2^{\prime}\left(\frac{1}{2}+\frac{\pi^2}{6{N}_r^2}\right)\right] $$
(59)
$$ {\eta}_c=\frac{d\left(2{N}_r\Gamma {h}_g{l}_r\right)}{dt}=2{N}_r\Gamma {h}_g{v}_r $$
(60)

where \( {k}_1^{\prime } \) and \( {k}_2^{\prime } \) are both coefficients. Γ is the glass material fracture (surface) energy. There should be a balance between the “input energy” (i.e., the increasing elastic energy) and the “output energy” (i.e., the fracture energy). Suppose that the proportion between the two parameters ηb/ηc, which is named as “energy conversion factor,” indicating the stability of the system or the ability for further crack growth, then the larger proportion would obviously lead to a greater degree of instability for the whole system. Thus for the system of PVB laminated plates in our study, the “energy conversion factor” ηb/ηc can be expressed as

$$ \frac{\eta_b}{\eta_c}=\frac{\pi {E}_g{h}_g}{2\Gamma {c}_g}\cdot {V}^2\cdot \frac{1}{v_r}\left\{\frac{1}{N_r}\left[{k}_1^{\prime }+{k}_2^{\prime}\left(\frac{1}{2}+\frac{\pi^2}{6{N}_r^2}\right)\right]\right\} $$
(61)

One may clearly see that the expression of the “energy conversion factor” contains three parts: (1) πEghg/2Γcg is determined by the material properties (here refers to the glass material), (2) V2 refers to the impact condition, and (3) rest of the expression f(vr, Nr) refers to the impact responses of the laminates. Thus the stability of the system is determined together by the material properties, impact condition, and the impact response of the material according to Eq. 61. Previous studies (Xu et al. 2011d; Chen et al. 2013) investigating the effect of the impact speed on the crack velocity have demonstrated that, with higher impact speed V that increasing the system instability ηb/ηc, both of the crack velocity vr and the radial crack number Nr tend to increase as well thus to lower the value of ηb/ηc such to maintain the balanced energy input and output in a large extent. Therefore, as the crack number Nr on the backing glass layer decreases under the same impact condition, the value of ηb/ηc would correspondingly increase due to the result of Eq. 61, which means the instability of the system is rising. Thus the system tends to response in a way to reduce this instability (i.e., the value of ηb/ηc) through propagating the cracks at higher speed, improving the value of vr in Eq. 61, which is consistent with our experiment results (Fig. 46).

The Relation Between the Crack Length and the Capability of Energy Absorption on the Impacted Glass Layer

Further, the total crack length for both two crack patterns on the impacted glass layer is calculated to study the capability of energy absorption of laminated glass influenced by the radial crack number on the backing glass layer.

Figure 47 shows the crack morphology on the impacted glass layer where both of the two crack patterns (i.e., radial crack and circular crack) appear long after the cracking of the backing glass plate. For the radial cracks on the impacted layer that are indirectly caused by the stress concentration brought by each generated crack on the backing glass layer as initial flaw (Lee et al. 2007) (which is also proved by our experiment investigation), they are completely overlapped with the radial cracks on the backing layer (Fig. 47). Thus, it is obvious that the morphology as well as the number of radial cracks in the impacted glass layer depends completely on that of the backing layer. The radial cracks on the impacted glass initiates from those on the backing glass plate through the depth direction (Lee et al. 2007).

Fig. 47
figure 47

Selected sequence of images depicting the crack patterns on both glass layers at the loading speed Vd = 3.7 m/s, with PVB thickness h = 0.76 mm

Specimen in response to lower impact speed V = 2.42 m/s is chosen for 100 repeated experiments. As expected, the radial crack number on the backing glass plate varies from 15 to 65 in these experiments. The total crack length for both two patterns on the impacted glass layer is calculated with different crack numbers on the backing glass layer (Fig. 48). Lr and Lc are defined as the total radial and circular crack length on the impacted glass layer, respectively. Thus, (Lr + Lc) refers to the total crack length on the impacted glass plate. The radial crack length linearly increases with crack number on the backing glass layer due to the uniform growth of all cracks while the variation trend of circular crack length is decreasing (Fig. 48). Circular cracks always initiate long after the completed propagation of radial cracks on the impacted glass (Xu et al. 2010) (Fig. 47). Thus, before the initiation of circular cracks, the laminated plates can be regarded as Nr triangular beams with thickness 2hg and length l0 (which can be estimated as the side length of specimen) since the radial cracks on the two glass plates are completely overlapped (Fig. 47). The bending energy Ub0 of the plates is

$$ {U}_{b0}=4\pi {E}_g{h}_g^3{w}_0^2\left(1+{\pi}^2/3{N}_r^2\right)/\left(3{l}_0^2\right) $$
(62)
Fig. 48
figure 48

Crack length on the inner glass layer versus radial crack number (V = 2.42 m/s)

In the case with less radial crack number Nr, the bending energy would increase, resulting in stronger stress concentration in the radial direction, which makes the later generation of circular cracks easier. In addition, taking the law of energy conservation into account, the total fracture energy on the whole laminated plates can be estimated as 2ΓLhg, where L is the total crack length on the two glass plates. Thus, the fracture energy released on two glass plates can be estimated by calculating the crack length on each plate. For the plates with shorter radial crack length, leaving more residual energy, it will benefit the generation of circular cracks as well.

Although the increase of circular crack number contributes positively to the energy absorption of the laminated plates, the effect is much less compared with the radial cracks (Fig. 48). As the radial crack morphology on the impacted glass is directly determined by that on the backing glass layer, it is safe to conclude that the backing glass plate is the key that decides the fracture morphology and further the capability of energy absorbing of the whole laminated structure such that the material selection and structure design for the backing layer is extremely important for impact protection.

Summary

In this chapter, we have investigated the propagation characteristics for both radial and circular cracks in PVB laminated glasses by theoretical constitutive equations analysis, numerical simulation, experiments, and tests of impact.

Firstly, a damage-modified nonlinear viscoelastic constitutive relations model of PVB laminated windshield is suggested based on updated Lagrangian method to study the crack evolutions on windshield plate subject to pedestrian head.

Then, the constitutive relations are implemented into FEA software to simulate the pedestrian head impact with vehicle windshield after verified by a classical example and classical Hertzian pressure calculation model. Based on the FE model above, the internal stress of PVB laminated glass is analyzed and found that shear stress, compressive stress, and tensile stress are main causes of plastic deformation, radial cracks, and circumferential cracks, respectively for the laminated glass subject to impactor. By a parametric study based on the FE model above, the increase of Poisson’s ratio will lead the increase of crack angle. Impact velocity has a great influence of crack formations as well as the injuries of pedestrian head. Comparing with other numerical techniques, XFEM is arguably more efficient since it does not require remeshing during crack propagation. So the extended finite element method (XFEM) is adopted to numerically investigate the multiple crack propagation of PVB laminated glass subject to quasi-static indentation or low-speed impact. It is found that the critical accident information, such as the impact speed or damage stress, can be deduced from the crack pattern characteristics. A qualitative bridge can be established between numerical simulation result and real-world accident via the crack growth mechanism.

Lastly, for the experimental analysis of PVB laminated glass, crack branching based on crack fractal theory is investigated and an explicit expression describing the crack velocity and number of crack branching is proposed under quasi-static Split Hopkinson Pressure Bar (SHPB) compression experiments. For dynamic out-of-plane loading, we carried out preliminarily impact fracture experiments firstly to investigate the radial crack propagation behavior of PVB laminated glass subjected to light-weight impact by using high-speed photography. The time histories of the averaged radial crack tip position, propagation velocity, and acceleration are obtained. It is found that the steady-state cracking speed of PVB laminated glass is lower than that of pure glass, and it also increases with higher impactor speed and mass. It is also found that the supported glass layer would always initiate before the loaded layer and the final morphologies of radial cracks on both sides are completely overlapped even if they propagate at different time. Here, two different mechanisms of crack propagation on different glass layers are explained: the cracks on supported glass layer are dominantly motivated by the in-plane crack-tip stress such that a relatively longer time is needed for its propagation while for the cracks appearing much later on loaded layer, which are caused by the stress concentrated in depth direction for each generated crack on supported layer, the propagation time needed is largely reduced. Then, the further experiment investigating the radial and circular crack propagation of PVB laminated glass thoroughly is carried out. The parametric study on two dominant factors, i.e., impact velocity and PVB thickness is discussed aiming at the crack propagation of PVB laminated glass. Results show that cracking speeds of both radial crack and circular crack increase with higher impact velocity. Radial crack appears earlier than circular one and it propagates at higher speeds. Based on the mechanism of crack initiation and propagation, a semiphysical model describing the relationship between the maximum cracking velocity and influential factors is established. In addition, the Weibull statistical model is suggested considering various factors to describe the macroscopic crack pattern thus providing a theoretical evidence for engineering practice. Finally, the relation between radial crack velocity and crack numbers on the backing glass layer and the relation between the crack length and the capability of energy absorption on the impacted glass layer are investigated. Results show that the radial crack velocity on the backing glass layer decreases with the crack number under the same impact conditions during large quantities of repeated experiments. Thus, the “energy conversion factor” is suggested to elucidate the physical relation between the cracking number and the crack propagation speed. However, the fracture energy of the whole laminated plates is mainly determined by the radial crack length, which means that the radial crack number generated first on the backing glass plate is the key that decides the capability of energy absorbing of the whole laminated structure.