82
Temperature effects on the low velocity impact response of laminated glass with different types of interlayer materials Xiaowen Zhang a , Idris K. Mohammed b , Mengyao Zheng a , Nan Wu a , Iman Mohagheghian b,c , Guanli Zhang a , Yue Yan a , John P. Dear b a Beijing Institute of Aeronautical Materials, AECC, Beijing Engineering Research Centre of Advanced Structural Transparencies for the Modern Traffic System, Beijing, 100095, China b Department of Mechanical Engineering, Imperial College London, South Kensington Campus, London, SW7 2AZ, United Kingdom c Department of Mechanical Engineering Sciences, Corresponding author. E-mail address: [email protected] Corresponding author. E-mail address: [email protected] 1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 1 2 3 4

spiral.imperial.ac.uk · Web viewThe impact tests were conducted at an impact energy of 15J, which corresponds to an impact velocity of 3.83 m/s, and at four different temperatures

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Temperature effects on the low velocity impact response of laminated glass with different types of interlayer materials

Xiaowen Zhanga, Idris K. Mohammedb, Mengyao Zhenga, Nan Wua,

Iman Mohagheghianb,c, Guanli Zhanga, Yue Yana[footnoteRef:2], John P. Dearb[footnoteRef:3] [2: Corresponding author. E-mail address: [email protected]] [3: Corresponding author. E-mail address: [email protected]]

a Beijing Institute of Aeronautical Materials, AECC, Beijing Engineering Research Centre of Advanced Structural Transparencies for the Modern Traffic System, Beijing, 100095, China

b Department of Mechanical Engineering, Imperial College London, South Kensington Campus, London, SW7 2AZ, United Kingdom

c Department of Mechanical Engineering Sciences, University of Surrey, Guildford, Surrey, GU2 7XH, United Kingdom

Abstract

This paper investigates the influence of the interlayer material on the low velocity impact performance of laminated glass. The effect of temperature (50°C, 23°C, 0°C and -30°C) has been explored to observe damage mechanisms and the associated impact resistant properties of the laminated glass. The four interlayer materials investigated were: SGP–Ionoplast as employed in Sentry Glas® Plus, TPU- Thermo-plastic polyurethane, PVB-Polyvinyl butyral and a TPU/SGP/TPU hybrid interlayer. The impact resistance was measured in terms of load peak, absorbed energy, ultimate deformation and crack patterns. The low velocity impact results indicated that both the type of the interlayer materials and testing temperature have great influence on the impact resistant properties of the laminated glass. The laminated glass with SGP interlayer exhibited best impact resistant properties amongst the four structures at room temperature. However, as the temperature was varied, the TPU/SGP/TPU hybrid interlayer performed the best over the entire range of temperatures tested, which can better ensure the safety of the occupants in the vehicle. This is because the elastic and viscous properties of the interlayer materials greatly changes with the temperature caused by the different glass transition temperatures of the interlayer materials.

Keywords: laminated glass, impact properties, interlayer material, temperature

1. Introduction

Laminated glass has been widely used in many engineering applications such as aircraft and automobile windshields as well as architectural and security glazing due to its excellent safety properties. Laminated glass normally consists of two or more layers of glass plates which are bonded together with a transparent, thermoplastic elastomeric interlayer. Monolithic glass has good structural strength, but can fail dramatically when subjected to crash or impact. However, the polymer interlayer can maintain the structural integrity of the component after breakage of the glass and protect the passengers from flying fragments. Additionally, the polymer interlayer is responsible for absorbing the remaining impact energy after glass breakage. This is a vital design consideration especially in transport industry to ensure the safety of vehicle occupants.

Nowadays, several polymer interlayers have been used for the laminated glass including polyvinyl butyral (PVB), thermo-plastic polyurethane (TPU), ethylene-vinyl acetate (EVA) and ionoplast Sentry Glas® Plus (SGP). PVB, TPU and EVA perform as the soft interlayer materials, which have lower stiffnesses at room temperature; compared to SGP, which has shown a significantly stiffer response at room temperature (e.g. the elastic modulus of SGP is two orders of magnitude greater than PVB [1-3]). The structural performance of a laminated glass is significantly influenced by the properties of the interlayer material [3-5]. El-Shami et al. [4] investigated the stress development in laminated glass with different interlayer materials including regular polyvinyl butyral (PVB) as well as a stiffer formulation of PVB and found that the laminated glass with the stiffer formulation of PVB resulted in a significantly larger load resistance. PVB has been widely used as a general interlayer material in the car and architecture laminated glass field and is of much research interest [6, 7]. SGP has the advantage of higher mechanical strength and higher modulus and is effective in improving the bending strength of laminated glass. As a result, SGP is the recommended polymer interlayer for anti-hurricane architectural glazing[8, 9]. Another interlayer of interest is TPU which has good adhesion to inorganic glass and polymeric plates, good low temperature flexibility and good anti-aging properties [10, 11].

Impact performance of laminated glass is of great importance for transport applications and has attracted much research interest [12-16]. The failure behaviour of laminated glass subjected to dynamic impact varies greatly from monolithic glass, due to the contribution of both elastic properties and energy-absorption capacity of the interlayer [15, 17]. Low and high velocity impact performance of the laminated glass have been studied using different experimental methods, such as drop weight impact machine, pendulum impactor and gas-gun tests. Mohagheghian et al. [4, 5] performed low and high velocity impact tests on the laminated glass using different materials and types of constructions. The effects of interlayer thickness, type of glass and polymer type and multi-layering the polymer interlayer on the structural performance have been investigated. It has been concluded that increasing the polymer interlayer thickness in the laminated glass configurations has little effect on the stiffness of the structure, but the type of polymer interlayer has significant effect on the stiffness and strength of laminated glass plates. Chen et al. [18, 19] conducted a drop-weight test in combination with high-speed photography to investigate the in-plane crack propagation behavior in laminated glass plates. Velocity-time history curves along with the force-time history curves of both radial and circular crack were calculated and analyzed. It showed that the cracks on impacted plate appeared long after the full growth of those on backing plate. The cracks on the backing layer were motivated by in-plane stress concentration while the cracks on impacted layer were caused by stress concentration in depth direction.

The impact resistance performance of the laminated glass was determined by many factors, with most research focusing on the influence of glass type and thickness, interlayer type and thickness, type of impactor, boundary conditions, impact velocity and dimensions of the glass plates [20-22]. Zhang et al. [21] conducted laboratory tests and numerical simulations on the vulnerability of laminated glass windows subjected to windborne wooden block impact. Different weights of wooden debris and different thickness PVB interlayer were investigated. It has been found that interlayer thickness plays a dominating role in the penetration resistance capacity of the laminated glass windows subjected to windborne debris impact.

As can be understood from above, there is an extensive range of studies available in the literatures considering the impact performance of laminated glass windows. These studies mainly focused on laminated glass using conventional polymer interlayer materials tested at the room temperature, especially PVB as it is widely used in automotive and construction industries. However, firstly for the high-tech applications such as laminated windshield glass of aircraft and high-speed rail, the laminated glass windows can experience large temperature changing. It can range between -50°C in the high altitude flight and 70°C during its taking off and landing process for the aircraft use. For the high speed rail laminated glass, it also suffers from the high and low temperature application when used in the tropical and gelid regions. This temperature variation greatly affects the properties of polymer interlayer material and its suitability especially in terms of impact performance of the whole laminated glass structure [23].

In this paper, we aim to investigate the low velocity impact performance of laminated glass windows using various polymer interlayer systems. The focus will be on the effect of test temperature on the perforation resistance as well as damage development and fracture patterns of laminated glass windows employing various polymer interlayer systems. At temperature controllable chamber will be used for this purpose allowing tests at temperatures between -30°C and 50°C. The force-displacement, energy absorption and loading response of the laminated glass were monitored and analysed during the impact.

2. Experiments

2.1 Materials

The glass plates used in this study were manufactured using silicate float glass. The monolithic glass plates were manufactured in Beijing Institute of Aeronautical Materials (BIAM) with the dimensions of 100×150 mm and a thickness of 2.0mm. Three types of polymer interlayer material were employed: Ionoplast Sentry Glas® Plus (SGP) (from DuPont), Thermoplastic Polyurethane (TPU)-PF2300 (from PPG company) and Polyvinyl Butyral (PVB) (from DuPont). The lamination was conducted in an autoclave (in BIAM) by placing one layer (interlayer with single material) or three layers (multi-material interlayer) of polymer between two layers of glass. For lamination always the tin side of the glass was used, which was cleaned prior to lamination using analytical reagent ethanol. The lamination was conducted under the following processing conditions: the temperature and the pressure were raised to 137 °C and 10 bar respectively in 30 mins, then both were held for 210mins before lowering the temperature at a rate of 2.5 °C/min to 49°C and releasing the pressure naturally. Four different laminated glass configurations were manufactured. The details of each configuration are listed in Table 1, for each type of material (SGP, TPU, PVB – Case 1, 3 and 4); including a hybrid multilayer interlayer (TPU/SGP/TPU - Case 2) for comparison. Here, Case 1 has been taken as a reference case. It should be noted that thickness of Case 2, which has multi-interlayer, is different from the other cases here. The intention here is to see the effect of adding two thin layers of TPU on either side of SGP. The influence of the polymer interlayer thickness on the flexural stiffness and low velocity impact performance of laminated glass plates has been investigated previously (ref 3 and 22) and showed to have negligible effect.

Table 1 The interlayer configurations for samples investigated in this study.

Configuration

Glass and polymer layers

Total plate thickness (mm)

Case 1

2.0 mm glass/1.52 mm SGP/2.0 mm glass

5.52

Case 2

2.0 mm glass/0.38mm TPU/1.52mm SGP/0.38mm TPU/2.0 mm glass

6.28

Case 3

2.0 mm glass/1.52 mm TPU/2.0 mm glass

5.52

Case 4

2.0 mm glass/1.52 mm PVB/2.0 mm glass

5.52

2.2 Measurement and characterization

2.2.1 Low velocity impact tests

The impact tests were performed with an Instron CEAST 9350 drop weight equipped with a temperature-controllable chamber. The picture of the machine and the relative parts are shown in Fig. 1. Liquid nitrogen was used to lower the temperature in the chamber, for the low temperature testing conducted at 0°C and -30°C. The impactor used for the impact tests was hemispherical in shape with a diameter of 12.7 mm. The total dropping mass is approximately 2.048 kg including the carriage mass of 1.3 kg and the impactor mass of 0.748 kg. To measure the impact force history, a piezoelectric force transducer is mounted on the bottom side of a steel cross-bar. The laminated glass specimens were fixed inside the chamber using the clamping system which consists of four cuffs to properly fasten the specimen. The cuffs are made of steel and with a rubber cap at the head of the cuffs. The clamps have a minimum holding capacity of 1100 N. In this way it can help to prevent any plate movement during the impact test and also to ensure the same boundary conditions for all the experiments. The tips of the clamps are covered with the rubber caps with a durometer of Shore A 70-80. The cut-out in the plate is 75±1mm by 125±1 mm here. Three guiding pins are located such that the specimen shall be centrally positioned over the cut-out. The fixture shall be aligned to a rigid base using bolts or clamps. The schematic of clamping system is shown in Fig. 2[24].

For this machine, an anti-rebound system was also used to prevent a secondary impact on the laminated glass specimens after the first impact. The instrument is equipped with aninfrared sensor, whose position is adjustable by means of a measuring system. This sensor measures the displacement. Before starting the test the sensor must be set, depending on the position of the impactor. The data of time and force can be obtained from the machine directly. According to the ASTM D7136[24], the impact velocity (vi) can be obtained according to the data of the displacement sensor using the Eq. 1

Where vi =impact velocity, m/s; W12 = distance between leading edges of the first (lower) and second (upper) flag prongs, m; t1= time first (lower) flag prong passes detector; t2= time second (upper) flag prong passes detector; ti= time of initial contact obtained by the machine, s.

The time and force data obtained from the machine used to calculate the velocity of the impactor during the contact with the structure as follow:

Where v(t) is the impactor velocity at time t (t = 0 is the time when the impactor initially contacts the specimen), F(t) is the measured impactor contact force at time t, g is the gravity acceleration constant equal to 9.81 m s-2 and m is the total mass of the impactor. It should be noted that the vi is the initial velocity upon impact and can be calculated using equation 1. Subsequently, the impactor displacement as a function of time can be calculated as:

Where di is the impactor displacement from reference location (here is the surface of the specimens) at time t = 0. Based on equations 2 and 3, the absorbed energy transferred to the structure, Ea, during impact can be calculated by:

Fig. 1 Experimental setup used for low velocity impact tests with the Instron CEAST 9350 drop tower and the temperature-controllable chamber.

Fig. 2 The schematic of the clamping system[24]

The impact tests were conducted at an impact energy of 15J, which corresponds to an impact velocity of 3.83 m/s, and at four different temperatures of -30°C, 0°C, 23°C (i.e. room temperature) and 50°C. All the specimens were kept in the chamber for 20mins to achieve the desired equilibrium temperature before the impact was conducted (except at 23°C). The thickness of specimen was measured using a micrometer before the testing and the data was input into the system, thus the initial displacement (di) of the impactor from the specimens can be calculated automatically by the system. If the impact energy is higher than the largest gravitational potential energy of the impactor up to the limitation of the machine, an initial velocity will be produced by the system, but 15J is not a higher energy than the largest gravitational potential energy of the impactor, so the initial velocity of the impactor is zero here. Three specimens of each configuration were impacted at each temperature to ensure repeatability of the test results.

2.2.2 Crack patterns analyses

Crack patterns of the laminated glass specimens after impact were obtained using a digital camera. Since both glass layers were broken after impact it was hard to distinguish between the crack pattern in the front and rear glass plates. In order to take the picture of the fracture pattern in the front glass layer, the rear glass layer was painted in black prior to take pictures. This is firstly, to mask the cracks occurred in the rear glass layer and secondly, to make the cracks on the front impacted face more visible by creating a larger contrast using a black background. The black paint finish was obtained by dissolving black dye in alcohol, the black paint was then poured onto the surface of the specimens instead of spraying. In this way, the black paint can infiltrate the cracks reducing the reflection of the cracks. Another series of photographs were also taken to capture the fracture pattern in the rear glass layer. This time the fracture pattern in the front glass layer was masked by black paint. Although a series of other samples were used to capture the fracture pattern in either sides of the glass layer, the shape and density of the crack were quite repeatable. It should be noted that three repeated tests were conducted for each configuration at the particular temperature. Take the energy versus time curves as the typical representation, the repeated tests for each configurations are plotted together at room temperature to show the repeatability of the method, which will be shown in Section 3.1 Fig 6.

2.2.3 Dynamic mechanical analysis (DMA)

Dynamic Mechanical Analysis (DMA) was used to obtain the storage modulus (E’) and loss modulus (E’’) of the three interlayers. The tests were conducted in tension mode using a TA Instruments Q800 DMA machine at the frequency of 1 Hz with oscillation amplitude of 15 μm. The specimens had a rectangular geometry with the width of 6.17 mm. The free length of each specimen was measured individually after fixing the specimen in the clamp. The experiments were conducted under temperature sweep testing mode, the specimens were left at -100°C for 10 min to achieve thermal equilibrium, before heating up to 80°C with a heating rate of 2°C min-1.

When the interlayer materials are subjected to an oscillating load, there will be a phase hysteresis between the applied stress and the measured strain. This is caused by the hysteresis response of the viscous component in the material. As a result, the modulus of the material can be expressed as Eq. 2:

where E’ and E’’ are the storage and loss modulus respectively. The ratio of the storage over the loss modulus (E’/E’’) is defined as tangent of the phase angle (tan delta). In the present research, the temperature corresponds to the peak value of tan delta is considered as the glass tan delta transition temperature (Tg). It should be noted that the glass transition temperature measured here is for the testing frequency of 1 Hz. If the testing frequency changes, there will be a shift in the glass transition temperature for all three polymer interlayer materials.

3. Results

3.1 Different types of interlayer material

Typical force-deformation and energy-time curves obtained during impact testing of Case 1 (SGP) at room temperature (23°C) are shown in Fig. 3a and Fig. 3b respectively. The force versus deformation plot, as shown in Fig. 3a, can be divided into three stages: in Stage I, an approximately monotonic increase of the force over time until the first peak force, Ffl, at a deformation of, Dfl, which corresponds to the fracturing of the first layer of glass. Then the force decreases dramatically to almost zero, this phase is due to the compressibility of the interlayer material after the fracturing of the first layer of glass. In Stage II, the force increases over time until a second peak, Fsl, at a deformation of, Dsl, resulting in the fracturing of the second layer of glass. Beyond Fsl, the significant drop in force was due to crack propagation in the laminated glass. In Stage III, the deformation continues reaching the ultimate deformation, Dul, before the impactor rebounds due to the residual strength of the laminated glass after fracturing. The response in this region is mainly controlled by the strength of polymer interlayer.

The energy versus time plot (Fig. 3b) shows the initial impact energy that the machine imposed on the specimen with a value of approximately 15J. The absorbed energy (Ea) can be defined as the value of the plateau part of the energy curve, after the initial impact. The elastic energy (Ee) is then given by the difference between the initial impact energy and absorbed energy [25], corresponding to the rebound energy on the impactor after striking caused by the rebound of the laminated glass. However, in the case when the impactor completely perforates through the laminated glass plate, the peak energy will never reach 15 J. In this situation, the impactor carries a residual velocity, vr, and consequently residual kinetic energy of Er= account for the remaining energy. Small portion of energy might also be dissipated through friction and bending of the intact remaining parts as the impactor continues its movement.

Fig. 3 Typical impact response curves for laminated glass Case 1 (SGP), for an impact energy of 15 J at room temperature: (a) Force versus deformation curve, (b) Energy versus time plot.

In Fig. 4, both force versus deformation and energy versus time curves for all the studied laminated glass specimens with various types of interlayer materials are presented. All the experiments have been conducted at a test temperature of 23°C and an impact energy of 15 J (Cases 1 to 4). The values for absorbed energy (Ea), elastic energy (Ee), force and deformation for first glass layer failure (Ffl and Dfl), force and deformation for second glass layer failure (Fsl and Dsl) and the ultimate deformation (Dul) are reported in Table 2 respectively.

As shown in Fig. 4c and Table 2, Case 1 (SGP) showed the highest elastic energy followed by Case 2 (TPU/SGP/TPU) (with only slightly lower elastic energy). Case 3 (TPU) and Case 4 (PVB) showed very low elastic energies. The rebounding of impactors was greatly influenced by the viscoelastic properties of the interlayer materials. Viscoelasticity in the polymer interlayers involves both “elastic” and “viscous” contributions. The “elastic” part is responsible to the deformation recovery while its “viscous” part leads to a delay effect in strain recovery upon impact unloading [26]. Therefore, the above results were caused by the different mechanical properties of the interlayer materials, SGP possessing the highest modulus and yield strength at room temperature compared to the TPU and PVB interlayers [1, 2, 10]. This can be clearly seen in the DMA results in Fig. 5a where at room temperature the storage modulus (E’) of SGP is highest amongst the three interlayer materials used. The superior mechanical properties of SGP at room temperature also led to benefits in the higher strength, stiffness and creep resistance of the whole laminated glass, both before and after glass breakage [27].

All the above mechanisms contribute to the loading response during the impact are shown in Fig 4a. The initial part of the force-deformation curve is cropped in Fig. 4b to make the curves more distinct for the reader. The gradient of the force-deformation curves between Ffl and Dfl are different from each other after the force is higher than about 1 KN, with Case 1 (SGP) being higher than that of Case 2, and then Case 3 (TPU) and Case 4 (PVB), indicating that the higher modulus of the interlayer material resulted in a higher elastic bending strength of the laminated glass specimens (as shown in Fig 5a). It should be noted that, despite the different interlayer materials, the response in all four Cases was the same up to approximately 1 kN, which corresponded to the duration of the impactor being in contact with the glass before the impact force reached the interlayer material. For Case 2 with a hybrid multilayer interlayer (TPU-SGP-TPU), the gradient of the force-deformation curves between Ffl and Dfl over 1kN was, as expected, smaller than that of Case 1 but higher than that of Case 3. It might be speculated that this result was caused by the larger thickness of Case 2 interlayer, but it has been reported that the interlayer thickness effects on laminated glass unit structural response were minimal[3, 22]. Therefore, the true reason should be Case 2 used one layer of TPU as a transition layer between SGP and glass, with the shear transfer between glass and SGP being affected by TPU interlayers, which contributed to the lower Ffl value in Case 2 compared to that of Case 1.

After the fracture of both the glass layers, the residual loading must be carried by the interlayer materials. The lower ultimate deformation (Dul) of the specimen means higher loading carrying capacity and also better rebound capacity of the interlayer materials. Accordingly, Case 1 had the best loading carrying capacity after fracture of the glass plates, indicated by the smallest Dul value in Table 2 and the highest reverse velocity of the impactor in the velocity-time curve in Fig. 4d. The load carrying capacity of Case 2 was slightly less than that of Case 1, but its Dul value was almost half that of Case 3 and Case 4, which indicated that the addition of an SGP layer lead to an improved loading carrying performance of the laminated glass after impact. The results are coincident with the higher elastic modulus of SGP at room temperature as indicated by the DMA results in Figure 5.

Fig. 4 Effect of interlayer structure on the impact properties of laminated glass at 23°C (SGP – Case 1, TPU/SGP/TPU - Case 2, TPU - Case 3, PVB - Case 4), for an impact energy of 15 J: (a) Force - displacement over full 26 mm, (b) Force-displacement over first 6 mm,(c) Energy – time, (d) Velocity -time curves.

Fig. 5 The dynamic mechanical analysis results of: a) Storage Modulus (E’) and b) Tan Delta, plotted against temperature for SGP, TPU and PVB interlayers.

Table 2 Low velocity impact of laminated glass with different interlayers (SGP – Case1, TPU/SGP/TPU - Case 2, TPU - Case 3, PVB - Case 4) at 23°C.

Specimen

Ea(J)

Ee(J)

Ffl(kN)

Fsl(kN)

Dfl(mm)

Dsl(mm)

Dul(mm)

Case 1

11.57±0.12

3.31±0.13

4.98±0.46

4.46±0.51

1.13±0.05

2.34±0.23

10.04±0.2

Case 2

12.06±0.052

2.82±0.079

2.56±0.16

4.54±0.29

1.16±0.23

3.55±0.12

10.88±0.21

Case 3

14.01±0.85

0.99±0.85

2.10±0.24

2.70±0.43

1.03±0.24

3.32±0.46

20.69±1.53

Case 4

14.09±0.20

1.00±0.23

2.32±0.31

2.76±0.22

1.06±0.28

3.02±0.17

25.80±0.78

To illustrate the repeatability of the method, here, the energy versus time curves of the repeated tests for each configuration at room temperature are plotted together as shown in Fig. 6. It can be seen that the scatter of the same three specimens is not large, the method and analysis are considered to be reliable.

Fig. 6 The energy versus time curves of the repeated tests for each configuration at room temperature. SGP – Case 1, TPU/SGP/TPU - Case 2, TPU - Case 3, PVB - Case 4

3.2 Effect of temperature

The experiments were conducted on each of the four Cases at temperatures of 50°C, 23°C, 0°C and -30°C with an impact energy of 15 J. The force versus deformation and energy versus time curves for Cases 1-4 are presented in Fig. 7 and 8 respectively. For clarity, the initial part of the force versus deformation responses is plotted in Fig. 9.

Fig. 7 Force versus deformation responses for laminated glass plates with different interlayers: Case 1 – SGP, Case 2 – SGP/TPU/SGP, Case 3 – TPU, Case 4 – PVB tested at temperatures of 50°C, 23°C, 0°C and -30°C, with impact energy of 15 J.

Fig. 8 Energy versus time graphs for laminated glass plates with different interlayers: Case 1 – SGP, Case 2 – SGP/TPU/SGP, Case 3 – TPU, Case 4 – PVB tested at temperatures of 50°C, 23°C, 0°C and -30°C,with impact energy of 15 J.

Fig. 9 The initial part of the force versus deformation responses for laminated glass plates with different interlayers: Case 1 – SGP, Case 2 – SGP/TPU/SGP, Case 3 – TPU, Case 4 – PVB tested at temperatures of 50°C, 23°C, 0°C and -30°C, with impact energy of 15 J.

3.2.1 Case 1 (SGP Interlayer)

It can be seen from Table 3 and Fig. 7a that although the Ee value obtained at 0°C was the highest amongst the temperatures tested, the loading values of Ffl and Fsl were slightly higher at 23°C compared to 0°C. The higher stiffness of the SGP interlayer at lower temperatures provides a more rigid support against the localised deformation of frontal glass layer when comes to the contact with the impactor. As a result, the fracture in the glass layers occurs earlier. On the other hand, when both glass layers are broken the remaining load will be carried by the polymer interlayer itself. The higher stiffness of the SGP interlayer at low temperatures results in a greater elastic energy stored in the structure.

From Fig. 8a, it can be seen that for Case 1 the absorbed energy, Ea, value decreased as the temperature was lowered except at -30°C where Ea was almost equal to the impact energy. As can be seen in Fig. 5a, the stiffness of SGP interlayer increases as the temperatures are lowered. The stiffer interlayer material contributed to the higher strength and better elastic properties of the whole laminated glass specimens. However, if the temperature decreases to a certain level, as in the case of -30°C, the SGP interlayer material became brittle and susceptible to fracture instead of elastic deformation. As can be seen in Figs 7a and 9a, the laminated glass plate with SGP interlayer fractures at much lower loads and displacements. This results in much smaller elastic deformation stored in the glass layers prior to fracture. In contrast most of the absorbed energy was dissipated through fracture in the glass layers as well as deformation and subsequently fracture in the SGP interlayer. As the impactor perforate the structure in this case, the transferred energy to the structure does not reach 15 J as indicated in Fig 8a. The velocity of the impactor after strike is measured to be 1.13±0.21m/s (as shown in Fig. 10) and the corresponding residual energy of 1.34±0.47J. As mentioned earlier, the remaining energy which is only small portion of the initial energy, 0.39±0.37J, might be dissipated through friction or bending the remaining intact parts of the plate. The ultimate deformation, Dul, of the specimen after striking at -30°C could not be determined because there was no rebound of the impactor as shown in Fig. 7a.

The above results show that the monolithic SGP loses its ductility at low temperatures. This can lead to fracture in the polymer interlayer itself which consequently affects the structural integrity of the whole laminated glass. As post impact breakage of laminated glass is an important safety aspect in the design, usage of monolithic SGP is not recommended for low temperatures.

Fig. 10 Velocity versus time results of three repeat tests for Case 1 impacted at temperature of -30°C.

Table 3. The low velocity impact properties of the laminated glass at different temperatures with SGP interlayer.

T(°C)

Ea(J)

Ee(J)

Ffl(kN)

Fsl(kN)

Dfl(mm)

Dsl(mm)

Dul(mm)

50

14.28±0.39

0.72±0.36

1.53±0.25

1.94±0.18

0.88±0.12

2.37±0.13

18.25±0.26

23

11.57±0.12

3.31±0.12

4.98±0.46

4.46±0.51

1.13±0.053

2.34±0.23

10.04±0.15

0

10.14±0.23

4.86±0.23

3.83±0.37

3.14±0.25

1.26±0.048

2.82±0.088

10.21±1.33

-30

13.25±0.85

0.39±0.37

3.94±0.39

3.35±0.36

1.22±0.014

2.41±0.023

-

3.2.2 Case 2 (TPU/SGP/TPU Hybrid Interlayer)

From Fig. 7b and 9b it can be seen that as the temperature was decreased, the initial gradient of the force versus deformation curves was increased. The increase in stiffness of the laminated glass is because of increase in shear modulus of TPU as well as SGP interlayer at low temperatures. The higher the shear modulus the higher shear stresses transfer between the two glass layers and the higher the stiffness of the laminated glass.

The energy dissipation property of Case 2 at -30°C was better that of Case 1 (pure SGP) which showed poor elastic properties at -30°C. After the introduction of the TPU interlayer as the transition layer in Case 2, the impact resistance property of the whole structure was significantly improved because the TPU interlayer is much more flexible than SGP at -30°C as can be seen in the storage modulus (E’) versus temperature plots of Fig. 5a. This helps to retain the structural integrity of the laminated glass plates after both glass layers breakage and results in much better post breakage performance. From the DMA data shown in Fig 5b, it could be seen that the glass transition temperature, Tg, of TPU is -32°C, much lower than that of SGP (54°C) and close to the test temperature of -30°C. As a result, TPU layer has not yet reached its glassy state. No sign of fracture was observed in the SGP layer for this case because the TPU transition layer can dissipate some absorbed energy and the energy left to SGP become less, and will not cause the fracture of the SGP interlayer. Therefore, in this case SGP keep carrying the remaining load while TPU interlayer ensure the structural integrity of the laminated glass, which resulted in better impact property of the whole laminated glass.

Table 4. The low velocity impact properties of the laminated glass at different temperatures with TPU/SGP/TPU hybrid interlayer.

T(°C)

Ea(J)

Ee(J)

Ffl(kN)

Fsl(kN)

Dfl(mm)

Dsl(mm)

Dul(mm)

50

14.86±0.062

0.14±0.062

1.37±0.16

2.93±0.19

1.23±0.08

3.45±0.12

24.47±2.36

23

12.06±0.047

2.82±0.078

2.56±0.16

4.54±0.29

1.16±0.23

3.55±0.12

10.88±0.21

0

10.97±0.27

4.03±0.27

4.30±0.22

3.44±0.26

1.30±0.13

2.92±0.33

11.45±1.19

-30

10.94±0.58

4.06±0.58

4.36±0.33

2.24±0.13

0.96±0.20

2.48±0.25

8.08±1.11

The impact energy is dissipated through three possible mechanisms: i) fracturing in glass layer and in the polymer interlayer (in certain conditions), ii) the deformation of polymer interlayer and iii) the delamination between the polymer interlayer and glass layers. The energy dissipated through the delamination is believed to be smaller compared with the other two mechanisms. For the impact tests conducted at 50°C, the deformation of the laminated glass plate was much larger than the results obtained at the three lower temperatures, as indicated by the high value of Dul in Table 4. Therefore, it seems that the concentration of radial and circumferential cracks on both the glass layers tested at 50°C was less than the glass layers tested at 23°C and 0°C, shown in Fig. 14b. At low temperatures, the higher stiffness of the polymer interlayer provided a greater support against the local deformation of the frontal glass layer when comes to the contact with impactor. This resulted in a highly localised fracture pattern, as can be seen in Fig 14b. In contrast, at higher temperatures the polymer interlayer became softer and more complaint as a result. Therefore, the deformation is distributed over the larger area and consequently the fracture occurs over a larger area.

3.2.3 Case 3 (TPU Interlayer)

From the results shown in Fig. 8c and Table 5 for Case 3, it can be seen that the Ea value decreased while Ee value increased as the temperature was lowered except at -30°C. This is consistent with mechanisms leading to enhanced elasticity of TPU material when the temperature is lowered from 50°C to 0°C, however, -30°C is nearly as low as the Tg value of the TPU material, at which temperature TPU became less ductile capable of undergoing large deformations, inducing the relative lower Ee value of the specimen at -30°C compare with that at 0 °C.

However, when struck at 50 °C, the Ee value is about 0.077±0.0028 as the absorbed energy is 14.85±0.065J and the final kinetic energy of the impactor are 0.15 J, 0.08J, 0.00 J as the velocity of the impactor for the three testing was 0.38 m/s, 0.28m/s and -0.02m/s, as indicated in Fig. 11. This is because the TPU interlayer became very soft at 50 °C and its load carrying capacity was significantly reduced after the fracture of the glasses, and very little rebound happened for the impactor. Consequently serious deformation was induced as shown in Fig. 12.

Fig. 11 Velocity versus time results of three repeat tests for Case 3 impacted at temperatures of 50°C.

Fig. 12 Fracture plate of Case 3 impacted at 50°C with large deformation

Compared to Case 1 and Case 2, with the stiffer SGP interlayer, monolithic TPU provided inferior impact performance at high temperatures. Conversely, at low temperatures, TPU can still maintain good flexibility and thus the impact resistant properties of the laminated glass with the monolithic TPU interlayer (Case 3) were better than that with monolithic SGP interlayer (Case 1), maintaining the structural integrity after impact. However, the impact properties were still inferior to that of the laminated glass with TPU/SGP/TPU hybrid interlayers (Case 2). This can be seen by comparing Tables 4 and 5 for recoverable elastic energy, Ee of 1.26 J for TPU compared with 4.06 J for hybrid interlayer and for first peak force, Ffl, of 4.36 kN for TPU compared with 2.32 kN for hybrid interlayer at -30°C. It can be concluded that the hybrid interlayer structure of Case 2 has integrated the advantages of both Case 1 and Case 3.

Table 5. The low velocity impact properties of the laminated glass at different temperatures with TPU interlayer (Case 3)

T(°C)

Ea(J)

Ee(J)

Ffl(kN)

Fsl(kN)

Dfl(mm)

Dsl(mm)

Dul(mm)

50

14.85±0.065

0.077±0.0028

2.41±0.40

3.22±0.11

2.11±0.15

3.84±0.05

-

23

14.01±0.85

0.99±0.85

2.10±0.24

2.70±0.43

1.03±0.24

3.32±0.46

20.69±1.53

0

12.48±0.15

2.52±0.15

4.95±0.66

5.64±0.31

1.35±0.05

1.92±0.09

12.11±0.18

-30

13.74±0.10

1.26±0.10

2.80±0.18

2.64±0.54

1.06±0.02

2.46±0.12

13.00±1.64

3.2.4 Case 4 (PVB Interlayer)

Again, similar to Case 3, when struck at 50 °C, the energy-time curve was terminated before reaching the impact energy of 15J as shown in Fig 8d, for the similar reason as Case 3, as PVB interlayer became very soft at 50 °C and the specimen was almost penetrated during the impact, indicating that Case 4 also possessed poor impact resistant property at high temperature. The Ea value is approximately 14.05 J as seen in Table 6, indicating there still left some energy on the impactor after the striking, and the velocity of the impactor for the three experiments are 1.01m/s, 0.90m/s, 1.05m/s, then the Ee value is almost zero, as the same reason as that for Case 3. The Dul value is very large indicating that the rebound property of the PVB interlayer is very poor after the fracturing of the glass layers.

However, at 23°C and 0°C, the Case 4 laminated glass plate possessed similar impact resistant property to that of Case 3. The Ee values at 23°C and 0°C are relatively lower than that of Case 1 and Case 2 due to the high mechanical strength of the SGP material. However, the Dul value of Case 4 was the highest amongst all the structures at 23°C and 0°C, indicating that the elastic property and also the rebound capacity of the PVB interlayer was the lowest compared to the other three types of interlayer materials.

Also, Case 4 showed poor impact resistant property at -30°C which was fractured directly instead of being deformed upon being struck. This is because PVB had become brittle at -30°C which was lower than its Tg temperature of about 27°C, as indicated in Fig. 5b. Similar to the testing at -30°C of Case 1, the elastic energy in the glass plate can be obtained through the subtraction of the kinetic energy of the impactor from the difference between the initial impact energy (15 J) and the absorbed energy. As the velocity of the impactor after striking is still about 0.37±0.21m/s and the corresponding kinetic energy is about 0.17±0.16J, then the elastic energy in the glass plate is calculated to be about 0.86±0.67J.

Table 6 The low velocity impact properties of the laminated glass at different temperatures with PVB interlayer (Case 4)

T(°C)

Ea(J)

Ee(J)

Ffl(kN)

Fsl(kN)

Dfl(mm)

Dsl(mm)

Dul(mm)

50

14.05±0.59

0.00±0.00

1.26±0.23

1.50±0.30

0.58±0.15

1.26±0.23

-

23

14.09±0.20

1.01±0.23

2.32±0.31

2.76±0.22

1.06±0.28

3.02±0.17

25.80±0.78

0

11.60±0.23

3.40±0.23

3.16±0.16

2.64±0.23

1.16±0.28

2.83±0.05

19.18±1.13

-30

13.96±0.81

0.86±0.67

4.54±0.94

-

0.94±0.051

-

-

Fig. 13 Absorbed energy (Ea) and elastic energy (Eb) versus temperature curves for the four configurations tested at different temperatures with impact energy of 15 J.

In order to directly compare the impact performance of the different configurations, the absorbed energy (Ea) and elastic energy (Eb) versus temperature curves have been plotted in Fig. 13a and b. It can be seen that Case 2 (with TPU-SGP-TPU hybrid interlayer) has the lowest Ea value but highest Ee value at -30°C. Although Case 1 shows lower Ea value than Case 3 and Case 4, the laminated glass specimens are apt to be fracturing instead of deformation, leading to the lower Ee value. At 0°C, the sequence of Ea values for the four configurations is Case 1< Case 2< Case 4< Case 3, and contrast to the Ee values. This is also true for the results obtained at 23°C except the Ea value of Case 4 almost being same to Case 3. Lower Ea value is considered to be beneficial to ensuring the safety of vehicle occupants. For the high temperature impact, the Ea and Ee values of the four configurations are not very different from each other, but Case 3 and Case 4 are almost penetrated after the impact. Case 1 and Case 2 with at least one layer of SGP interlayer can keep the structural integrity of the whole laminated glass after impact.

3.3 Crack patterns

Images of the crack patterns of each Case tested at temperatures of 50°C, 23°C, 0°C and -30°C with an impact energy of 15 J are shown in Figure 14.

31

Fig. 14 Crack patterns on the front (f) and back (b) face of the laminated glass for a) Case 1 – SGP, b) Case 2 – SGP/TPU/SGP, c) Case 3 – TPU, d) Case 4 – PVB tested at temperatures of 50°C, 23°C, 0°C and -30°C with impact energy of 15 J.

3.3.1 Case1-SGP interlayer

As the temperature of the test was lowered, the density of the radial and circumferential cracks decreased except for the specimen impacted at -30°C, which resulted in brittle fracture as shown in Fig. 14a. Also, the damage becomes more localized because the rigid SGP interlayer is bad for the energy transmittance between the two glass layers. This was consistent with the lowered energy absorption results shown in Fig. 8a, and fewer circumferential cracks were visible on the back layer of the glass impacted at 0°C. This was because the SGP became stiffer at lower temperatures and consequently the laminate bending resistance increased, as mentioned in Section 3.2.1. However, if the test temperature is dropped below a certain level, e.g. tests at -30°C, ductility of the SGP would be affected and resulted in fracture in the SGP layer upon impact. This undermines the benefit from the increased bending resistance of the structure at low temperatures and results in lower absorbed energy values.

3.3.2 Case 2-TPU/SGP/TPU interlayer

For the impact test conducted at 50°C, the deformation of the glass layers and the interlayer material was the largest of the four temperatures tested for Case 2 according to Table 4. From Fig. 14b, it can be seen that the density of radial and circumferential cracks on both the glass layers tested at 50°C was less than the glass layers tested at 23°C, 0°C and -30°C. Again, the damage became more localized as the temperature dropped also caused by the bad energy transmittance capacity of the rigid TPU/SGP/TPU interlayer at lower temperatures.

3.3.3 Case 3-TPU interlayer

From Fig. 12, 14c and Table 5, it can be seen that there was large deformation and fragmentation of the specimen when impacted at 50°C. The radial and circumferential cracks on both the glass layers became sparse as the temperature was lowered to 0°C, this is because the absorbed energy at 0°C is the lowest.

3.3.4 Case 4- PVB interlayer

The fractured specimen with the PVB interlayer is shown in Fig. 14d. It was very different from that of Case 3 with TPU as the interlayer material. The specimen impacted at 0°C resulted in the least deformation and fragment sizes, corresponding to a lower Ea value. Similar to Case 3, high temperature caused larger deformation and fragmentation, while the interlayer was almost fully penetrated due to the poor load carrying capacity of the PVB material after fracture of the glass layers at 50°C. However, at the very low temperature of -30°C, the PVB interlayer became brittle, leading to direct fracture of the specimen when impact occurred, similar to what happened to Case 1 at -30°C.

4. Conclusions

In this paper, the low velocity hard impact properties and damage mechanisms of laminated glass with four different types of interlayer materials were evaluated. All the structures were tested at four different temperatures whilst the impact energy was kept fixed. Through a systematic study, the effect of the interlayer material and temperature on the energy dissipation, loading distribution and crack patterns were measured and analysed. The following conclusions were drawn:

· The type of interlayer has a strong effect on impact performance of laminated glass plates. The stiffest polymer interlayer, SGP (Case 1), provided the highest elastic bending strength to the laminated glass, resulting in the best load carrying capacity after the fracturing of the glass layers at room and high temperatures. However, the SGP became brittle when the temperature was lowered to -30°C, leading to the direct fracturing upon impact.

· The laminated glasses using soft polymer interlayer TPU (Case 3) offers better impact performance at lower temperatures compared to laminates using SGP interlayer. This is because of the good flexibility of the TPU at low temperatures. However, it is almost penetrated at higher temperatures, e.g. 50°C as a result of relatively low stiffness of TPU (i.e. 50°C is well above the glass transition temperature of the material at which the stiffness is significantly reduced), which is considered to be bad for the occupants and instruments in the cabin.

· The laminated glass using PVB (Case 4) as an interlayer, showed similar impact performance to TPU (Case 3) at 50°C, 23°C and 0°C and almost perforation when impacted at 50°C. Unlike the impact performance of Case 3 at -30°C, the laminated glass using PVB interlayer fractured directly after striking, which was similar to the SGP interlayer specimen at -30°C.

· The hybrid interlayer (Case 2), consisting of an SGP layer sandwiched between two layers of TPU, showed similar impact performance to monolithic SGP (Case 1) at 50°C, 23°C and 0°C, but at -30°C its impact resistance properties were considerably better than the other three configurations because of the lowest absorbed energy and less damage to the specimens, which is considered to be safer for the occupants and instruments in the cabin. It can be concluded that the hybrid interlayer (Case 2) possessed the most attractive impact resistant properties among the four configurations studied over the entire range of temperatures tests, with the hybrid interlayer integrating the advantages of both SGP and TPU.

Acknowledgement

This research was performed in collaboration with the Imperial research teams at the BIAM Centre for Materials Characterisation, Processing and Modelling at Imperial College London. This research work is financially supported by the Innovation Foundation Programme of Beijing.

References

[1] P. A. Hooper, B. R. K. Blackman, J. P. Dear. The mechanical behaviour of poly (vinyl butyral) at different strain magnitudes and strain rates. Journal of Materials Science, 2012; 47: 3564–3576.

[2] X. Zhang, Y. Shi, H. Hao, J. Cui. The mechanical properties of ionoplast interlayer material at high strain rates. Materials & Design, 2015, 83, 387-399.

[3] I. Mohagheghian, Y. Wang, L. Jiang, X. Zhang, X. Guo, Y. Yan, A. J. Kinloch, J. P. Dear. Quasi-static bending and low velocity impact performance of monolithic and laminated glass windows employing chemically strengthened glass, European Journal of Mechanics A/Solids, 63, 2017, 165-186.

[4] M. M. El-Shami, S. Norville, Y. E. Ibrahim. Stress analysis of laminated glass with different interlayer materials. Alexandria Engineering Journal, 2012, 51, 61-67.

[5] I. Mohagheghian, Y. Wang, J. Zhou, L. Yu, X. Guo, Y. Yan, M. N. Charalambides, J. P. Dear. Deformation and damage mechanisms of laminated glass windows subjected to high velocity soft impact. International Journal of Solids and Structures, 2017, 109, 46-62.

[6] J. Xu, Y. Li, B. Liu, M. Zhu, D. Ge. Experimental study on mechanical behavior of PVB laminated glass under quasi-static and dynamic loadings, Composites: Part B, 2011, 42, 302-308.

[7] M. S. Shetty, L. R. Dharani, J. Wei, D. S. Stutts. Failure probability of laminated architectural glazing due to combined loading of wind and debris impact. Engineering Failure Analysis, 2014, 36, 226-242.

[8] J. Belis, J. Depauw, D. Callewaert. Failure mechanisms and residual capacity of annealed glass/SGP laminated beams at room temperature. Engineering Failure Analysis, 2009, 16, 1866-1875.

[9] M. Overend, C. Butchart, H. Lambert, M. Prassas. The mechanical performance of laminated hybrid-glass units. Composite Structures. 2014. 110, 163-173.

[10] L. Zhang, X. Yao, S. Zang, Q. Han. Temperature and strain rate dependent tensile behavior of a transparent polyurethane interlayer. Materials and Design, 2015, 65, 1181-1188.

[11] G. Shim, S. Kim, D. Ahn, J. Park, D. Jin, D. Chung, S. Choi. Experimental and numerical evaluation of transparent bulletproof material for enhanced impact energy absorption using strengthened-glass/polymer composite. Composites Part B, 2016, 97, 150-161.

[12] J. Xu, Y. Li, X. Chen, Y. Yan, D. Ge, M. Zhu, B. Liu. Characteristics of windshield cracking upon low-speed impact: Numerical simulation based on the extended finite element method. Computational Materials Science, 2010, 48, 582-588.

[13] S. Zhao, L. R. Dharani, L. Chai, S. D. Barbat. Analysis of damage in laminated automotive glazing subjected to simulated head impact. Engineering Failure Analysis, 2006, 13, 582-597.

[14] J. Xu, Y. Li, B. Liu, M. Zhu, D. Ge. Experimental study on mechanical behavior of PVB laminated glass under quasi-static and dynamic loadings. Composites: Part B, 2011, 42, 302-308.

[15] A. Ball, H. W. Mc Kenzie. On the low-velocity impact behavior of glass plates. J De Physique IV, 1994; 4(C8): 783–788.

[16] T. Pyttel, H. Liebertz, J. Cai. Failure criterion for laminated glass under impact loading and its application in finite element simulation. International Journal of Impact Engineering, 2011, 38, 252-263.

[17] Y. M. Tsai, H. Kolsky. A study of fractures produced in glass blocks by impact. Journal of the Mechanics and Physics of Solids, 1967, 15(4): 263-278.

[18] J. J. Chen, J. Xu, X. F. Yao, B. H. Liu, X. Q. Xu, Y. M. Zhang, Y. B. Li. Experimental investigation on the radial and circular crack propagation of PVB laminated glass subject to dynamic out-of-plane loading. Engineering Fracture Mechanics, 2013, 112-113, 26-40.

[19] J. J. Chen, J. Xu, X. F. Yao. Different driving mechanisms of in-plane cracking on two brittle layers of laminated glass. International Journal of Impact Engineering, 2014, 69: 80-85.

[20] X. Zhang, H. Hao, Z. Wang. Experimental study of laminated glass window responses under impulsive and blast loading. International Journal of Impact Engineering, 2015, 78, 1-19.

[21] X. Zhang, H. Hao, G. Ma. Laboratory test and numerical simulation of laminated glass window vulnerability to debris impact. International Journal of Impact Engineering, 2013, 55, 49-62.

[22] R. A. Behr, J. E. Minor, M. P. Linden. Load duration and interlayer thickness effects on laminated glass. Journal of Structural Engineering. 1986, 112, 1441-1453.

[23] S. M. Walley, J. E. Field, P. W. Blair, A. J. Milford. The effect of temperature on the impact behaviour of glass/polycarbonate laminates. International Journal of Impact Engineering, 2004.30, 31-53.

[24] ASTM D7136M-07, USA. Standard Test Method for Measuring the Damage Resistance of a Fiber-Reinforced Polymer Matrix Composite to a Drop-Weight Impact Event.

[25] R. M. Boumbimba, M. Coulibaly, A. Khabouchi, G. Kinvi-Dossou, N. Bonfoh, P. Gerard. Glass fibers reinforced acrylic thermoplastic resin-based tri-block copolymers composites: Low velocity impact response at various temperatures. Composite Structures, 2017, 160, 939-951.

[26] B H. Liu, T. N. Xu, X. Q. Xu, Y. Wang, Y. T Sun, Y. B. Li. Energy absorption mechanism of polyvinyl butyral laminated windshield subjected to head impact: Experiment and numerical simulations. International Journal of Impact Engineering, 2016, 90, 26-36.

[27] S. J. Bennison, C. A. Smith, A. Van Duser, A. Jagota. Structural performance of laminated glass made with a “Stiff” interlayer. Glas. Process. Days, 2001, 283-287.

32

01234560246810Force [kN]Deformation [mm]02468101214160246810121416Energy [J]Time [ms]abImpact EnergyElastic EnergyAbsorbed EnergyFflFslDflDslDulStage IStage IIStage III

-2-1012340246810121416Velocity [m/s]Time [ms]Case 1Case 2Case 3Case 4-101234502468101214161820222426Force [kN]Deformation [mm]Case 1Case 2Case 3Case 4-10123450123456Force [kN]Deformation [mm]Case 1Case 2Case 3Case 402468101214160246810121416Energy [J]Time [ms]Case 1Case 2Case 3Case 4abcd

00.20.40.60.811.2-100-50050100Tan DeltaTemperature [C]SGPTPUPVB1.E+001.E+011.E+021.E+031.E+04-100-50050100E' [MPa]Temperature [C]SGPTPUPVBab

02468101214160246810121416Energy [J]Time [ms]50230-3002468101214160246810121416Energy [J]Time [ms]50230-3002468101214160246810121416Energy [J]Time [ms]50230-3002468101214160246810121416Energy [J]Time [ms]50230-30Case 1 -SGPCase 2 -HybridCase 3 -TPUCase 4 -PVBabcd