You are on page 1of 94

EXPERIMENTAL & NUMERICAL STUDY OF DYNAMICALLY LOADED BOLTED &

HYBRID (BOLTED/BONDED) JOINTS

By

Aiswarya Venkadachalam

A THESIS

Submitted to
Michigan State University
in partial fulfillment of the requirements
for the degree of

Engineering Mechanics - Master of Science

2014
ABSTRACT

EXPERIMENTAL & NUMERICAL STUDY OF DYNAMICALLY LOADED BOLTED &


HYBRID (BOLTED/BONDED) JOINTS

By

Aiswarya Venkadachalam

Joining of composite materials is commonly achieved by bolting, bonding, or a combination of

the two methods (hybrid) in engineering structures such as aircraft, marine and automotive. If not

designed properly, dynamic loading of these joints can be critical to the integrity of the structure,

as failure initiates at fastener-material interfaces. This demands a thorough understanding of the

structural response of joints subjected to various loading conditions. The present study aims at

predicting the strength of single-lap, bolted and hybrid joints with various geometric

configurations under impact loading conditions. Experiments are conducted in a Drop Weight

Impact Testing Machine (DWITM) where tensile loading is applied by a specially designed

fixture. Results show that at room temperature, the strength of the hybrid joints with shorter edge

distances is greater than or comparable to the strength of bolted joints with larger edge distances.

At elevated temperatures (60°C), there is a 25-35% reduction of strength for joints with smaller

edge distances while joint strength for large edge distances is invariant with respect to

environmental conditions. In addition, the effect of clamping force on bolted joint strength is

investigated using a Split Hopkinson Pressure Bar (SHPB). Results confirm that at higher

loading rates, bolting torque plays a dominant role on the bearing strength of the joint structure.

Enhanced 3D finite element (FE) models are developed to substantiate the experimental

predictions and results are found to be in good agreement.


I dedicate this thesis to my mom and my loving husband

iii
ACKNOWLEDGMENTS

I would like to sincerely acknowledge my thesis advisor, Dr. Srinivasan Arjun Tekalur,

for his continuous support throughout the research and writing process. I am always

grateful for his technical assistance, persistent guidance and encouragement which

certainly helped me on both my research as well as on my career. I also thank my

committee members, Dr. Dahsin Liu and Dr. Xinran Xiao for taking time to review my

thesis and making suggestions. In addition, special thanks to Dr. Abhishek Dutta, Dr.

Guojing Li and Dr. Ermias Gebrekidan Koricho for their invaluable advice and insightful

comments in many research aspects which undoubtedly improved the quality of the work.

I sincerely thank our Ph.D. students Wei Zang and Andrew Joe Vander Klok for their

contribution to this project. I am indebted to my colleague, David Gonzalez for helping

me directly and indirectly throughout my research. I thank him for all his help,

discussions, support and fun we had for the last one and a half years. I would like to thank

my team mates Wu Zhou, Steven Ellis Utz and Peter Noel Howes for their assistance in

carrying out experiments. Furthermore, I thank Mike McLean, Adam Klein, Todd Pasch,

Aida Montalvo and the late Gail Berry for their extended help. I gratefully acknowledge

the financial support provided by TARDEC.

Finally, I thank my parents, my better half and my friends, for always being a source of

love, inspiration and encouragement in all my needy times. Without you all this wouldn’t

have been possible.

iv
TABLE OF CONTENTS

LIST OF TABLES vii

LIST OF FIGURES viii

KEY TO SYMBOLS xiii

Chapter 1 Introduction 1
1.1 Joining methods and strength calculations 1
1.2 Testing methods 2
1.3 Bolt preload and joint behavior 3
REFERENCES 6

Chapter 2 Evaluating Bolted Joint Behavior at High Strain Rates 7


2.1 Abstract 7
2.2 Introduction 8
2.2.1 Failure modes on composite joints observed from static testing 9
2.3 Experimental study 11
2.3.1 Material and model description 11
2.3.1.1 Test setup and joint configuration 11
2.3.2 Results and discussion 14
2.3.2.1 Effect of bolt-pretension 15
2.4 Numerical study 19
2.4.1 Model description 19
2.4.2 ABAQUS/Standard – Bolt-Preload analysis 22
2.4.2.1 Contact Properties 22
2.4.2.2 FE Mesh 23
2.4.2.3 Bolt preload and boundary conditions 23
2.4.3 Results and discussion 26
2.4.3.1 Bolt pre-tension 26
2.4.3.2 FE model validation 27
2.4.3.3 Effect of slip from numerical model 31
2.4.3.4 FE results for tension test 34
2.5 Limitation 37
2.6 Conclusion 37
REFERENCES 39

Chapter 3 Evaluating Hybrid Joint Behavior at Intermediate Strain Rates 40


3.1 Abstract 40
3.2 Introduction 40
3.2.1 Joining methods 41
3.2.2 Testing methods 42
3.2.3 Failure modes on composite joints observed from static testing 43

v
3.3 Experimental work 44
3.3.1 Static Testing Method 44
3.3.2 Low velocity Impact Testing Method 44
3.3.3 Specimen configurations and materials 47
3.4 Experimental Results and Discussion 49
3.4.1 Experimental observations from static analysis 49
3.4.1.1 Response of bolted joints 49
3.4.1.2 Effect of e/d ratio 50
3.4.1.3 Response of bonded joints 52
3.4.1.4 Failure modes on bonded joints 53
3.4.1.5 Response of hybrid joints 54
3.4.1.6 Effect of e/d ratio 55
3.4.1.7 Effect of joint type 56
3.4.2 Experimental observations from impact analysis 58
3.4.2.1 Force equilibrium 58
3.4.2.2 Response of bolted joints 61
3.4.2.3 Effect of e/d ratio 61
3.4.2.4 Response of bonded joints 64
3.4.2.5 Failure modes on bonded joints 64
3.4.2.6 Response of hybrid joints 66
3.4.2.7 Effect of e/d ratio 67
3.4.2.8 Effect of joint type 68
3.4.2.9 Effect of loading rates 69
3.4.2.10 Effect of hole-clearance 70
3.4.2.11 Effect of environmental conditions 71
3.4.2.12 Effect of composite adherend 72
3.5 Conclusion 74
REFERENCES 77

Chapter 4 Conclusions and Recommendations 80


4.1 High strain rate impact testing 80
4.2 Static and low velocity impact testing 80
4.3 Recommendations 81

vi
LIST OF TABLES

Table 2-1 Metal-metal bolted joint specimen configuration and applied bolt preload. 13

Table 2-2 Bolted joint specimen configurations used for numerical analysis. 20

Table 2-3 Material properties of the specimen, bolt and the loading bars used in
numerical analysis. 21

Table 3-1 Material properties of S2 glass-SC 15 epoxy laminates [16]. 49

Table 3-2 Dimensions of pulse shaping materials used to achieve force equilibrium
within the specimen under impact loading. 59

vii
LIST OF FIGURES

Figure 1.1 Strain rates and its corresponding experimental techniques used for
material characterization 2

Figure 1.2 Behavior of joint under static loading 3

Figure 2.1 Typical failure modes observed in single-lap, single-bolt composite joints
loaded at the hole: (a) shear-out (b) cleavage (c) net tension and (d)
bearing. 10

Figure 2.2 Schematic diagram of Split Hopkinson compression bar used in current
study shown in (a), placement of the load cell to monitor bolt preload
shown in (b) and the effect of preload on bolt and contact surfaces shown
in (c). 12

Figure 2.3 Bolt joint configuration and key parameters associated with the current
study. 13

Figure 2.4 Plot of incident pulse, transmission pulse and reflected pulse obtained from
incidence and transmission bar. 15

Figure 2.5 Transmission force of aluminum/aluminum bolted joint clamped at 500lbf. 16

Figure 2.6 Variation of force to slip and maximum bearing force of


aluminum/aluminum bolted joint for varying preloads. 16

Figure 2.7 Transmission stresses of aluminum/aluminum bolted joint for varying


preloads: 0lbf, 100lbf, 300lbf, 500lbf and 700lbf. 17

Figure 2.8 Strain histories from the load cell outfitted on the bolt for varying preloads:
0lbf, 100lbf, 300lbf, 500lbf and 700lbf. 18

Figure 2.9 Loading rate vs. bolt preload of the specimen under high strain rate impact. 18

Figure 2.10 Finite element (FE) model of split Hopkinson pressure bar (top) used for
numerical analysis and schematic represent of applied bolt preload with
equivalent nut pressure on FE model (bottom). 21

Figure 2.11 Plastic properties of the aluminum sample used in numerical analysis. 22

Figure 2.12 Schematic represent of master and slave surfaces assigned to the finite
element model. 23

Figure 2.13 Finite element model showing mesh on aluminum/aluminum bolted joint. 24

Figure 2.14 Steps employed to apply bolt preload on finite element model in static
solver. 25
viii
Figure 2.15 Schematic represent of bolt partition where preload is applied on finite
element model. 25

Figure 2.16 Stresses developed in the bolt region after applying a preload of 500lbf
from finite element modelling. 26

Figure 2.17 Stresses developed on aluminum-bolt contact surfaces for a applied preload
of 500lbf. 26

Figure 2.18 Location of strain gauges in experimental method & corresponding nodal
positions in finite element model. 27

Figure 2.19 Comparison of transmission stresses of aluminum/aluminum bolted joint


from experimental and finite element results for varying preloads: 0lbf,
100lbf, 300lbf, 500lbf and 700lbf. 28

Figure 2.20 Comparison of bearing strength of aluminum/aluminum bolted joint from


experimental and finite element results for varying preloads: 0lbf, 100lbf,
300lbf, 500lbf and 700lbf. 29

Figure 2.21 Percentage of error in aluminum/aluminum bolt joint strength from finite
element predictions by comparing against experimental results. 30

Figure 2.22 Schematic represent of load cell location in experimental method &
corresponding section where nodal values are extracted from finite element
model. 30

Figure 2.23 Comparison of preloading effect on aluminum/aluminum bolted joint from


experimental and finite element results for varying preloads: 0lbf, 100lbf,
300lbf, 500lbf and 700lbf. 31

Figure 2.24 Schematic represent of contact zone on the metal constituents. 32

Figure 2.25 Different phases on slipping event in aluminum/aluminum bolted joint


preloaded at 500lbf subjected to high loading rates. 33

Figure 2.26 Representation of bolt-metal contact during the event of slipping (a) at
region I and (b) at region I and III. 33

Figure 2.27 Correlation of strain-time history and bolt slip on aluminum/aluminum


bolted joint preloaded at 500lbf from finite element predictions. 34

Figure 2.28 Variation of bearing strength of aluminum/aluminum bolted joint under


tension loading for different e/d ratios from finite element analysis. 35

Figure 2.29 Failure mode predictions of aluminum/aluminum bolted joint under tension
loading for different e/d ratios from finite element analysis. 35

ix
Figure 2.30 Variation of bearing strength of aluminum/aluminum bolted joint under
tension loading for different w/d ratios from finite element analysis. 36

Figure 2.31 Failure mode predictions of aluminum/aluminum bolted joint under tension
loading for different w/d ratios from finite element analysis. 36

Figure 3.1 Typical failure modes observed in single-lap, single-bolt composite joints
loaded at the hole: (a) shear-out (b) cleavage (c) net tension and (d) bearing
(Reproduced from Figure 2.1). 43

Figure 3.2 Drop tower test setup (left), fixture setup with holding specimen (middle)
and joint specimen (right). 45

Figure 3.3 Fixture used to convert compression to tension load in drop tower machine. 46

Figure 3.4 Hybrid joint specimen configuration used for impact analysis and
geometrical parameters: Front view (shown in left) and side view (shown in
right). 48

Figure 3.5 Bolted joint geometry used for impact analysis. 48

Figure 3.6 Aluminum/S2 glass-SC 15 hybrid samples placed on the specially designed
fixture to maintain alignment and constant pressure. 49

Figure 3.7 Load-displacement curve of Aluminum/S2 glass-SC 15 bolted joint with


edge distance 1d under static loading. 50

Figure 3.8 Failure loads of Aluminum/S2 glass-SC 15 bolted joints with different e/d
ratios under static loading. 51

Figure 3.9 Failure modes of Aluminum/S2 glass-SC 15 bolted joints with different e/d
ratios under static loading. 52

Figure 3.10 Load-displacement curve of Aluminum/S2 glass-SC 15 bonded joint with


bond area 0.5 in. x 1 in. under static loading (load bottom to top). 53

Figure 3.11 Typical failure modes definition of an adhesive bonded joint: (a) cohesive
failure, (b) adhesive failure, (c) mixed failure and 53

Figure 3.12 Failure modes of Aluminum/S2 glass-SC 15 bonded joints with bond area
0.5 in. x 1 in. under static loading. 54

Figure 3.13 Load-displacement curve of Aluminum/S2 glass-SC 15 hybrid joint with


edge distance 1d under static loading (load bottom to top). 54

Figure 3.14 Failure load predictions of the Aluminum/S2 glass-SC 15 hybrid joints
with different e/d ratios under static loading. 55

x
Figure 3.15 Failure modes of Aluminum/S2 glass-SC 15 hybrid joints with different e/d
ratios under static loading. 56

Figure 3.16 Failure modes of Aluminum/S2 glass-SC 15 hybrid joints with e/d =1 and
bond area of 0.5 in. x 1 in. under static loading. 56

Figure 3.17 Failure load comparison of Aluminum/S2 glass-SC 15 bolted, bonded &
hybrid joints under static loading 57

Figure 3.18 Force vs. displacement plot of Aluminum/S2 glass-SC 15 joints under
static loading (a) bolted joint with e/d = 1, (b) bonded joint with bond area
0.5 in. x 1 in. and (c) hybrid joint with e/d =1. 58

Figure 3.19 Force histories of dog-bone aluminum specimen from top and bottom load
cell on using different pulse shaping materials: (a) Buna-N-rubber, (b)
polyurethane, (c) polypropylene and (d) latex rubber. 60

Figure 3.20 Force history of Aluminum/S2 glass-SC 15 bolted joint with edge distance
1d under impact loading from top load cell (load right to left). 61

Figure 3.21 Bolt transition & rotation of Aluminum/S2 glass-SC 15 bolted joint with
edge distance 1d under impact loading (load right to left). 62

Figure 3.22 Failure loads of Aluminum/S2 glass-SC 15 bolted joints with different e/d
ratios under intermediate loading rates. 63

Figure 3.23 Failure modes of Aluminum/S2 glass-SC 15 bolted joints with different e/d
ratios under impact loading. 63

Figure 3.24 Force history of Aluminum/S2 glass-SC 15 hybrid joint with edge distance
1d under impact loading from top load cell (load right to left). 64

Figure 3.25 Schematic of joint slip (yellow line) and bending (blue line) of
Aluminum/S2 glass-SC 15 bonded joint with bond area 0.5 in. x 1 in. under
impact loading (load right to left). 65

Figure 3.26 Failure modes of Aluminum/S2 glass-SC 15 bonded joints with bond area
0.5in. x 1in. under impact loading. 65

Figure 3.27 Force history of Aluminum/S2 glass-SC 15 hybrid joint with e/d = 1 under
impact loading (load right to left). 66

Figure 3.28 Failure load predictions of the Aluminum/S2 glass-SC 15 hybrid joints
with different e/d ratios under impact loading. 67

Figure 3.29 Failure modes of Aluminum/S2 glass-SC 15 hybrid joints with different e/d
ratios under static loading. 68

xi
Figure 3.30 Failure load comparison of Aluminum/S2 glass-SC 15 bolted, bonded &
hybrid joints under impact loading. 69

Figure 3.31 Failure load comparison of Aluminum/S2 glass-SC 15 bolted, bonded &
hybrid joints under static, intermediate and high loading rates. 70

Figure 3.32 Behavior of Aluminum/S2 glass-SC 15 hybrid joints with e/d = 1 for
varying hole clearances under intermediate loading rates. 71

Figure 3.33 Failure load comparison of Aluminum/S2 glass-SC 15 bolted & hybrid
joints with e/d = 1 and 3 between room temperature and 60°C. 72

Figure 3.34 Bearing strength comparison between Aluminum/S2 glass-SC 15 and


Aluminum/E-glass bolted joints with e/d =1 and 4 under intermediate
loading rates. 73

Figure 3.35 Bearing strength comparison between Aluminum/S2 glass-SC 15 and


Aluminum/E-glass hybrid joints with e/d =1 and 4 under intermediate
loading rates. 73

Figure 3.36 Failure mode comparison between Aluminum/S2 glass-SC 15 and


Aluminum/E-glass bolted and hybrid with e/d = 1 and 4 under impact
loading rates. 74

xii
KEY TO SYMBOLS

xiii
Chapter 1 Introduction

1.1 Joining methods and strength calculations

In engineering structures, joining of two or more structural components is often achieved by

mechanical fastening or bonding. Mechanical fastening (uses bolts or rivets) is the common

method of joining structural parts due to their relative ease of installation and resistance to

environmental degradation. Bolted joints are mostly preferred for transferring high loads thereby

widely used in large scale structures. However, high stress concentration regions are developed

around the hole locations in bolted joints where crack can initiate to cause structural failure.

Bonded joint is the ideal choice for thin walled structures where fatigue is of prime concern,

mostly used in aircraft and marine structures. This method of joining offers high efficiency

through smooth load transference, but it has some drawbacks. The strength of the bonded joint

depends heavily on surface treatment of the adherend members and operating conditions of the

joints. Also, the cost of maintenance for bonded joints is much greater than that of bolted

configuration. The strength of the joint depends on the joining method and its maximum load

carrying capacity. Bolted joint strength can be calculated as,

where, bearing area is defined as the thickness of the joint times the diameter of the bolt-hole.

Bonding strength can be determined as,

1
1.2 Testing methods

Strain-rate is defined as the rate at which the deformation occurs within a material with respect to

time. Understanding the behavior of joints under different strain rates is essential to optimize the

design and installation of joints in any structural members.

Figure 1.1 Strain rates and its corresponding experimental techniques used for material
characterization

Figure 1.1 illustrates different types of strain rates and its corresponding experimental

techniques used for characterizing material behavior. Standard testing machines are available to

execute quasi-static analysis to predict the strength of the materials at very low strain rates (1s-1).

These testing methods are used to perform both compression and tension analysis. The

experimental investigation in intermediate strain rate ranges (10 0 s-1 to 102 s-1) employs drop

tower testing machine for compression loading condition [1]. To investigate the behavior of

joints at higher strain rates (102 s-1 to 104 s-1), Split Hopkinson Pressure Bar (SHPB) is generally

used for both compression and tension analysis. Most conventional tools for dynamic

investigations are used to characterize material behavior. However, it is extremely difficult to

study the structural response of joints in both SHPB and drop tower testing system.

2
1.3 Bolt preload and joint behavior

In bolted joints, the applied pre-torque on the bolt and nut causes tension in the fastener and

subsequent compression on the material constituents. This preload helps in holding the joining

members together without slipping and plays a predominant role in defining the frictional force

between the contact surfaces. The amount of torque (T) required to achieve desired preload

( can be calculated as,

where, K is the nut factor usually considered as 0.2 and D is the diameter of the bolt [2].

Figure 1.2 Behavior of joint under static loading

When a static load is applied on a clamped bolted joint, four distinct regions [3] are evident in

force-displacement response as shown in Figure 1.2. In region 1, as the applied load increases

the entire joint will undergo elastic deformation where there is no relative movement of the

joining members are visible until it reaches a critical load. The magnitude of the region 1 entirely

3
depends on the amount of applied bolt-torque. In region 2, the joint slips until bolt-hole clearance

is eliminated. Relative displacement between the matting surfaces is evident in this region. In

region 3, significant plastic deformation can be seen until it reaches the failure of the joint

(region 4).

4
REFERENCES

5
REFERENCES

[1] K. T. Ramesh, “High Rates and Impact Experiments,” in Springer Handbook of


Experimental Solid Mechanics, W. N. S. J. Prof, Ed. Springer US, 2008, pp. 929–960.

[2] Zhou S, Wang Z, Wu X and Zhou J, "Experimental and numerical investigation on bolted
composite joint made by vacuum assisted resin injection," Compos Part B: Eng, p.
45(1):1620–8, 2013.

[3] G. Kelly, “Quasi-static strength and fatigue life of hybrid (bonded/bolted) composite single-
lap joints,” Composite Structures, vol. 72, no. 1, pp. 119–129, Jan. 2006.

6
Chapter 2 Evaluating Bolted Joint Behavior at High Strain Rates

2.1 Abstract

Bolted joints are used extensively in automobile and aerospace applications. Though bolting is

considered to be an easy and non-destructive method of joining, these fasteners are often

considered to be detrimental under impact loading conditions. Therefore, it demands specific

attention to a thorough understanding of the mechanical response of bolted joints at high rates of

loading. In this study, enhanced three dimensional (3D) finite element (FE) models are

developed for single-lap, single-bolt, metal-metal joints in order to observe their structural

behavior when subjected to impact loading in a Split Hopkinson Pressure Bar (SHPB). FE

solutions are compared against experimental results and are found to be in good agreement for

compression loading in SHPB. The validated FE models are further used to investigate the

failure modes of bolted joints under tensile loading condition. The following issues are

thoroughly investigated from the numerical model: (a) the effect of joint slip (b) influence of

geometric parameters on failure mode and (c) fastener preloading effect. Two different

geometrical parameters are investigated: edge distance-to-bolt diameter (e/d) ratio chosen from 1

to 5 and width-to-bolt diameter ratio (w/d) ranging from 1 to 4. The intent of this paper is to

provide a relationship between bolt pre-tension, loading rate and mode of failure by

computational method for high loading rates. These observations from the FE model will provide

ample information to characterize the crashworthiness of bolted structures where impact loading

is of prime concern and can be used as an alternative tool for structural joint test in SHPB.

7
2.2 Introduction

Bolted joints have been significantly used in the automobile and aeronautical industries for many

years where joining two or more structural members together are required. This kind of fastening

technique serves as a leading method for joining structural members by considering its ease of

installation and subsequent disassembly for inspection, maintenance, and repair work when

compared to adhesive or hybrid joints (bolted/bonded). Furthermore, bolts are relatively

inexpensive which makes it easier for replacement. However, due to the structural discontinuities

in their geometry, stresses are concentrated around the bolt-hole locations in the joint assembly,

which makes this region susceptible to failure. These higher stress regions result in a reduction in

structural joint strength which in turn leads to damage and failure of the entire structure.

Therefore, to employ bolted joints more effectively and efficiently, a thorough understanding of

joint mechanisms is highly necessary. The structural response of mechanical joints is

substantially different for different loading conditions. In particular, the dynamic response of

mechanical joints is more complicated because of the development of highly concentrated stress

regions, localized non-linear stiffness, inertial effects, dynamic friction, and damping factors

which are more likely to adversely affect the load carrying capacity of the joint. In the past

several years, exhaustive studies have been performed to understand the mechanical behavior of

bolted joints under quasi-static loading [1], [2], [3] but only a few have addressed the impact

response of mechanical joints [4].

In order to design joints where safety is of paramount concern, detailed knowledge of the stress

distribution in the critical regions is essential especially concerning failure modes of the

assembly. The study of bolted joint behavior is mostly parametric. Its mechanical response

depends on non-dimensional geometric parameters such as w/d, l/d, d/t and e/d where w is the

8
width of the joint, l is the length of joint, t denotes thickness, e indicates the edge distance from

the center of the bolt hole, and d is the diameter of the bolt used for non-dimensionalization.

Previous works on bolted joints are mainly focused on predicting the joint load to failure for

different geometrical configurations and materials. It has been shown that under quasi-static
0 -1
loading, tested at a strain rate of 10 s , the maximum bearing stress before failure (both in

compression and in tension) forms an asymptotic region for varying geometrical parameters.

Mode of failure depends on loading rate and tends to converge after certain values of w/d and e/d

ratios have been exceeded (approximately 2 for composite joints) [1], [2], [3], [4]. Under impact
2 -1 4 -1
loading conditions, tested at a strain rate of (10 s to 10 s ), the maximum bearing stress of

the joints are significantly higher than the joints subjected to quasi-static loading and

convergence of the asymptotic region of failure mode shifts to e/d ratio 3 (approximately) for

composite joints [4]. Early attempts were made to experimentally study the effects of clamping

force using different sized washers [5]. Also, the effects of bolt pre-load on a composite bolted

joint subjected to compression loading have been investigated [6]. Observations from these

studies clearly indicate that the bolt pre-load varies with applied force within the bolt and can

also be used as a key parameter to determine bearing mode of damage.

2.2.1 Failure modes on composite joints observed from static testing

Figure 2.1 illustrates the typical failure modes of single-lap, single-bolt mechanical joints

frequently observed during experimental investigations: shear-out, cleavage, net tension and

bearing failure [2], [3]. Mixed failure modes (combination of shear-out, cleavage or net tension)

are mostly noticed in the joints subjected to high loading rates where edge distance to diameter

(e/d) ratio is less than 3 [4]. Bearing failure is predominant when e/d ratio is over 3. Since

bearing failure is more gradual and progressive, it is often considered to be the preferred failure

9
as it can withstand loads even after the initiation of failure. All the other failure modes are

catastrophic and detrimental, which has to be avoided in structural member.

Figure 2.1 Typical failure modes observed in single-lap, single-bolt composite joints loaded
at the hole: (a) shear-out (b) cleavage (c) net tension and (d) bearing.

Performing experiments on bolted joints are still an ongoing process, since experimental data are

considered to be the most reliable datum line even today. But, in recent development of

numerical methods, extensive finite element analysis can be possible which is time consuming

but inexpensive and can predict detailed stress-strain distributions within the joint & modes of

failure. But care should be taken that the finite element model and analysis should mimic exact

physical system and realistic loading conditions.

The objective of this paper is to develop an efficient finite element bolted joint model subjected

to impact rates of loading in SHPB and to study its structural response by numerical approach.

10
This paper also discusses the significance of the role of bolt pre-tension and joint slip in bearing

mode of failure for metal-metal bolted joints. Joint slip is the relative movement of the faying

surfaces where the applied axial load surpasses the frictional force produced by the clamping

force of the fastener. The maximum amount of joint slip is predicted as twice the size of hole-

clearance but an average of one-half size of hole-clearance is observed in practical events. In this

work, a relation between bolt-pretension load, loading rate and maximum force transfer prior to

slipping condition is investigated. Experiments are carried out on split Hopkinson compression

bar and the results are in good agreement with FE solutions which predicts the dependence of

joint strength on loading rate.

2.3 Experimental study

2.3.1 Material and model description

2.3.1.1 Test setup and joint configuration

Experimental set-up for impact testing on SHPB is schematically represented in Figure 2.2. The

SHPB consist of split-bars (incident & transmission) which are made-up of 6061 T6 Aluminum

material with stiffness of 70GPa and has an identical diameter and length of 0.75 in., and 6 feet

respectively. The smaller diameter bar helps in generating high strain rates to characterize the

dynamic behavior of the bolted joint. The incident bar, commonly known as loading bar, is

loaded by a compressive pulse on the impact of the striker launched by a gas gun. The stress

pulse passes through the incidence bar, and at bar-specimen interfacial surface a portion of the

wave gets transmitted through the specimen, and the other portion gets reflected back. Since the

loading rate of incident pulse plays a major role in investigating the material response, it is

important to maintain force equilibrium within the specimen. Therefore, pulse shapers are used

that elongates the pulse width which helps in attaining force equilibrium within the specimen.

11
Both the incident and transmitted pulses are recorded by the strain gages which are bonded on

the split-bars. Electrical-strain gages are installed on the mid-span of each split-bar and are

positioned in diametrically opposite side in order to cancel out bar bending. Strain gages are

connected to a half-bridge circuit configuration. The output response of the circuit is amplified

by an Ectron amplifier (model513-2A) and data acquisition is carried out by using digital

oscilloscope (LeCroy-model354A).

Figure 2.2 Schematic diagram of Split Hopkinson compression bar used in current study
shown in (a), placement of the load cell to monitor bolt preload shown in (b) and the effect
of preload on bolt and contact surfaces shown in (c).

12
Figure 2.3 Bolt joint configuration and key parameters associated with the current study.

In addition, at the end of the transmission bar, a soft material is placed to prevent the

transmission bar from overshooting. A single-lap, single-bolt, metal-metal (Al 6061 T6) bolted

joint is tested for varies preloading conditions. To study the preloading effect, joints are outfitted

with calibrated load cell as shown in Figure. Lubricant is applied at the bar-specimen interfaces.

It helps in reducing the bar-sample interfacial frictional effects on loading. The geometric

description of the tested sample is shown in Figure 2.3. Details on applied preloads and

specimen configurations are presented in Table 2-1. For all the tested specimens the diameter of

the bolt is 6.35 mm.

Table 2-1 Metal-metal bolted joint specimen configuration and applied bolt preload.

13
2.3.2 Results and discussion

In this study, experiments are carried out in split Hopkinson compression bar which has been

developed to obtain loading rates from 200MN/s to 900MN/s. Stress equilibrium within the

sample is verified by comparing force data from the incidence, reflected and the transmission

pulses according to the wave theory [7].The maximum bearing stress of the bolted joint is

calculated as follows:

Equation 1
- Cross-sectional area of the transmission bar

- Young’s Modulus of the transmission bar

- Axial strain of the transmission pulse

- Bolt-hole diameter

- Specimen thickness

Parametric study has been executed on all 5 cases of the specimen series S1to observe the strain-

rate behavior of the bolted joints. However, to avoid multiple overlaps, experimental data

acquired from the incident and transmitted bar for the specimen clamped at 500lbf are alone

shown in Figure 2.4 It can be seen that the width of the incidence pulse is 150μs and only half

the intensity of the incident pulse is transmitted through the specimen and the remaining is

reflected back to the incident bar.

14
Figure 2.4 Plot of incident pulse, transmission pulse and reflected pulse obtained from
incidence and transmission bar.

2.3.2.1 Effect of bolt-pretension

Torque tightening in the bolted assembly helps the adherend members in contact by creating

tension in the bolt. The main objectives of this experiment are: (a) to investigate the influence of

torque tightening on joint behavior and (b) to examine the change in joint strength subjected to

impact loading conditions. Since the w/d ratio and e/d ratio of the tested samples are already

chosen to be in the asymptotic region, joint failure is confined to bearing mode of failure.

According to the transmitted force data from the experimental observation, behavior of bolted

joints can be classified as four stages. (a) Initial loading, (b) slip, (c) post slip and (d) bearing

damage. Different stages are described in Figure 2.5.

Force to slip (start of slipping stage) and max bearing force (onset of bearing damage stage) from

the experiments are plotted against bolt pretension and it is presented in Figure 2.6. This plot

shows that there is no substantial difference in force to slip between 100lbf and 300lbf. But, it

increases significantly when the value of bolt pretension increases from 300lbf to 500lbf. As we
15
further increase the torque tightening, force to slip drops down. The similar trend is also

observed for max force to failure.

Figure 2.5 Transmission force of aluminum/aluminum bolted joint clamped at 500lbf.

Figure 2.6 Variation of force to slip and maximum bearing force of aluminum/aluminum
bolted joint for varying preloads.

However, the influence of torque can be more phenomenal in slipping and the strain on the bolt

itself. Figure 2.7 and Figure 2.8 shows the transmission stress data for all 5 preloading

condition and the strain history from the load cell outfitted in the bolt. With reference to the

Figure 2.7, it is evident that the slipping can be delayed when we increase the torque tightening

16
of the bolt. For the preloading condition of 100lbf, slip occurs at 40μs. Whereas for higher

preloads (500lbf), slip occurs at 47μs and subsequent increase in preloads doesn’t contribute

much in slipping phenomena. From the overall strain plot (Figure 2.8), it is evident that the

increase in clamping causes smaller deviation in strain except for 100lbf. The strain on the bolt

reduces considerably for higher preloads (700lbf).

Figure 2.7 Transmission stresses of aluminum/aluminum bolted joint for varying preloads:
0lbf, 100lbf, 300lbf, 500lbf and 700lbf.

The loading rate experienced in the joint assembly for varying clamping conditions is presented

in Figure 2.9. The plot shows that the loading rate is higher for hand tightened bolted joints and

it decreases as we further increase the clamping force on the bolt. This is because, when the

clamping force in the bolt increases, it increase the compression on the metal adherends which

in-turn increase the frictional force between the faying surfaces. This helps in preventing the

relative displacement of the adherend members. Provided the bolt-hole clearance is maintained

as constant for all tested samples with the decrease in relative velocity, loading rate is expected

17
to decrease. But for the specific geometrical configuration of tested samples, it is observed that

the specimens pre-loaded at 300lbf are subjected to less loading rate.

Figure 2.8 Strain histories from the load cell outfitted on the bolt for varying preloads:
0lbf, 100lbf, 300lbf, 500lbf and 700lbf.

Figure 2.9 Loading rate vs. bolt preload of the specimen under high strain rate impact.

18
2.4 Numerical study

The mechanical behavior of bolted joint is complex under high strain rate loadings. Therefore, it

is important to substantiate the experimental results by computational results. A three

dimensional finite element model is developed in ABAQUS 6.11.2. Numerical analysis is carried

out for specimen series S1 for different preloading condition subjected to compression loading as

listed in Table 2-1. The model is further employed to study the response of bolted joint for

tensile loading as listed in Table 2-2. In order to observe the effect of slip, the bolt-hole

clearance is specified to the model as 0.1778 mm (0.007 inch), which is considered as the

standard hole-clearance for a bolt diameter of 6.35 mm (0.25 inch). Joint assembly is developed

with the entire SHPB setup in order to replicate the exact loading scenario as shown in

. Option to apply bolt pretension is currently not available in ABAQUS/Explicit solver.

Therefore, ABAQUS/Standard solver (static) is used to apply bolt preload in the assembly.

Clamped bolted joint is then imported to ABAQUS/Explicit solver in order to analyze the joint

behavior using non-linear dynamic formulation by setting-up the nodal positions from static

solution as the initial state for current analysis.

2.4.1 Model description

Geometric characteristics of the assembly and SHPB are imitated as described in the prior

explanations. In this investigation, the main scope is to model Split Hopkinson Pressure Bar

(Figure 2.10) to generate the exact compressive loading state irrespective of the computational

difficulties encountered in finite element analysis. Therefore, some modifications on geometry of

the SHPB model and bolt is incorporated which is not having a significant role in response of the

complete assembly in the event of loading. Changes encompassed in FE model are explained in

detail as follows: On SHPB model, striker is eliminated, since load is applied on surface of

19
incidence bar as a pressure wave. This helps in avoiding wave-splitting occurs at the interfacial

surface between incident bar & striker. Stopper is included as a rigid body at the end of

transmission bar. Nut and Bolt are modeled as an integral part and their threads are also

neglected (this assumption has been used by other researchers also).

Mechanical properties of the specimen (Aluminum) and bolt (Steel) are shown in the Table 2-3.

Sample is prescribed with elastic-plastic formulation (Figure 2.11) whereas bolt and loading bar

(Aluminum) are specified only with elastic property which are not the key part of investigation.

Densities of the materials are defined beforehand in static analysis which will be used in

subsequent dynamic analysis.

Table 2-2 Bolted joint specimen configurations used for numerical analysis.

20
Figure 2.10 Finite element (FE) model of split Hopkinson pressure bar (top) used for
numerical analysis and schematic represent of applied bolt preload with equivalent nut
pressure on FE model (bottom).

Table 2-3 Material properties of the specimen, bolt and the loading bars used in numerical
analysis.

21
Figure 2.11 Plastic properties of the aluminum sample used in numerical analysis.

2.4.2 ABAQUS/Standard – Bolt-Preload analysis

2.4.2.1 Contact Properties

Contact interactions are developed using surface-surface based discretization method by

enforcing linear penalty formulation with finite sliding. Master surface are assigned to more stiff

materials to avoid penetration into the slave surface as shown in Figure 2.12. Four interactions

properties are prescribed to the model: Metal - Metal, Bolt - Metal, Nut – Metal and Bar

(Incidence & Transmission) – Metal interactions. Frictional coefficient for interaction between

bolt-metal & nut-metal is prescribed as 0.7 & 0.3 respectively. Normal contact behavior is

imposed on the model with hard pressure over-closure relationship. Also, tangential contact

behavior is also prescribed by allowing finite sliding formulation. Frictionless contact is defined

for bar-specimen interfacial surface. To avoid complications in convergence of contact solution,

adjustment of slave surfaces are allowed to be exactly in contact with the master surfaces only

for few contact pairs. In others, adjustments of slave surfaces are not permitted. Contact

properties used are mostly based on the study of McCarthy et al. [8].

22
Figure 2.12 Schematic represent of master and slave surfaces assigned to the finite element
model.

2.4.2.2 FE Mesh

Finite element mesh of bolted joint assembly is shown in Figure 2.13. 8-noded linear elements

(C3D8R) are used to mesh the bolted joint assembly along with reduced-integration and

hourglass control. In order to get accurate solutions, a fine mesh is created around the bolt-hole

surfaces and in the regions of loading bars where strain gauges are located. A total of 12,444

elements are created for the metal surfaces and 6764 elements are constructed for the bolt part.

2.4.2.3 Bolt preload and boundary conditions

Bolt pretension is applied to the assembly using ABAQUS/Standard solver. Bold preload

can be calculated as,

Where, is the bolt-torque applied, K is the nut factor usually considered as 0.2 and D is the

diameter of the bolt. As shown in Figure 2.14 two steps approach is used for applying bolt

pretension on the sample which is influenced from the study of Song Zhou et al. [9].

23
Figure 2.13 Finite element model showing mesh on aluminum/aluminum bolted joint.

In the first step, bolt preload (0lbf – 700lbf) is applied to the assembly by clamping the free ends

of the metals (only in transverse directions). A “Partition surface” is created on the bolt along

bolt-diameter, where bolt pretension is applied normal to its surface. The partition of bolt cross

section where clamping force is applied is shown in Figure 2.15. By pulling the bottom portion

of elements in upward direction & top portion of elements in downward direction, ABAQUS can

automatically generate internal stresses which correspond to the applied load, thereby changing

the length of the bolt. Since bolt & nut are modeled as an integral part, equivalent pressure load

is applied on the nut surface to generate symmetrical stress distribution within the bolt cross

section along the axis (in order to mimic the exact loading condition). In the second step, length

of the bolt is fixed by holding every other constraints from step one. Time-dependent loading is

applied to the FE model using ABAQUS/Explicit solver. Stress pulses are applied at the end of

the incident bar by allowing the bar displacement along axial direction.

24
Figure 2.14 Steps employed to apply bolt preload on finite element model in static solver.

Figure 2.15 Schematic represent of bolt partition where preload is applied on finite element
model.

25
2.4.3 Results and discussion

2.4.3.1 Bolt pre-tension

Figure 2.16 shows the distribution of through-thickness stresses along the bolt shank pre-loaded

at 500lbf. As it can be seen, the shank of the bolt is subjected to a uniform tensile stress and

subsequent compressive stresses are observed in the metal, nut and bolt head region. Also,

significant stress concentrations are noticed on the edge of the bolt-hole region on the metal

surfaces caused by the bearing of the bolt as shown in Figure 2.17. The total compression

produced by the pretension of the bolt ranges from 6.5 to 96 MPa.

Figure 2.16 Stresses developed in the bolt region after applying a preload of 500lbf from
finite element modelling.

Figure 2.17 Stresses developed on aluminum-bolt contact surfaces for a applied preload of
500lbf.

26
2.4.3.2 FE model validation

To the best of authors’ knowledge, this study is the first effort to develop a SHPB model for

structural analysis. It implies the importance of validating the accuracy of the model before

further investigation. In this session, comparison of experimental data (strain gauges reading) is

carried out against the numerical results (axial stress values) for the compression loading in

SHPB. In FE model, stresses are extracted and averaged from two diametrically opposite nodes

on the transmission bar. The location of the strain gauges (experiment) and corresponding nodal

positions (FEA) are shown in the Figure 2.18.

Figure 2.18 Location of strain gauges in experimental method & corresponding nodal
positions in finite element model.

FE analysis is carried out for a total time period of 1000μs with a time step of 1μs. Data are

extracted and averaged for each time step. The respective transmission stresses for specimen

series S1 are plotted against computational results and are shown in Figure 2.19. It can be seen

that the behavior of FE model are comprehensively in good agreement with the experimental

results. The effect of slipping is not much evident from the FE transmission data. Potential

reasons could be: (a) time step in analysis and (b) applied nut-pressure.

In the current analysis, time step is maintained at 1μs. If we further bring down the time

increment, the possibility of visualizing the slipping effect in transmission pulse could be higher.

27
In contrary, it will be an additional increase in computational cost. Also, the slipping phenomena

can be investigated in detail from the specimen itself. Therefore it is not obligatory to refine the

time step of the analysis further.

Figure 2.19 Comparison of transmission stresses of aluminum/aluminum bolted joint from


experimental and finite element results for varying preloads: 0lbf, 100lbf, 300lbf, 500lbf
and 700lbf.

28
In ABAQUS model, an equivalent nut-pressure is applied to the joint assembly for all preloading

condition in order to have a uniform stress distribution along the through thickness direction of

the bolt. To predict the preload strain experienced in the bolt from the numerical model, applying

pressure on the nut surface is unavoidable.

The FE model is hence validated by comparing the maximum bearing strength (Equation 1) of

the bolted joint from the experimental data. The comparison of bearing strengths for all 5

clamping conditions and FE error percentage are shown in Figure 2.20 and Figure 2.21

respectively. With reference to Figure 2.20, it is evident that bearing strength of the bolted joint

increases with increase in preload up to 500lbf and decreases for subsequent preload.

Figure 2.20 Comparison of bearing strength of aluminum/aluminum bolted joint from


experimental and finite element results for varying preloads: 0lbf, 100lbf, 300lbf, 500lbf
and 700lbf.

Figure 2.21 shows that the FE solutions are in very good agreement with experimental results

for higher preloading conditions. The model is also used to verify the behavior of bolt itself.

Hence, the comparison of the bolt strain data is carried out amongst the load cell information and

29
average nodal values (nominal strain) across the bolt cross section as shown in Figure 2.22. The

respective transmission stresses for specimen series S1 are plotted against computational results

and are shown in Figure 2.23. This study is executed only to compare the behavior of the bolt

and but not to compare the exact strain values because of the difference in positions of load cell

and nodal locations where data are extracted. From Figure 2.23, it can be seen that the behavior

of bolt in FE model for higher preloading conditions (300lbf, 500lbf and 700lbf) are in good

compliance with the experimental results.

Figure 2.21 Percentage of error in aluminum/aluminum bolt joint strength from finite
element predictions by comparing against experimental results.

Figure 2.22 Schematic represent of load cell location in experimental method &
corresponding section where nodal values are extracted from finite element model.

30
Figure 2.23 Comparison of preloading effect on aluminum/aluminum bolted joint from
experimental and finite element results for varying preloads: 0lbf, 100lbf, 300lbf, 500lbf
and 700lbf.

2.4.3.3 Effect of slip from numerical model

Though the effect of slipping is not so evident from the transmission bar, it can be studied in

detail from the model itself. All the information presented here in this session is for joint

assembly clamped at 500lbf. Contact region of the bolted joint is shown in Figure 2.24.

31
From the FE model, slipping event can be classified as two phases: (a) Preliminary slip and (b)

Secondary slip (Figure 2.25). Before slipping occurs, there is no relative displacement between

the faying surfaces as the loading pulse just reaches the joint assembly at 315μs which is referred

as initial loading. In the first phase of slip, the pulse reaches halfway through the specimen (at

350μs). It initiates the movement of the contact surfaces between the metals. The load

transference is entirely by the friction between the metal surfaces but still there is no contact

between the bolt and the adherend. This phase extends till the adherend metal displaces to make

contact with the bolt surface.

Figure 2.24 Schematic represent of contact zone on the metal constituents.

In the second phase, the load transference by the friction exceeds and bolt-metal contact is

achieved at the bay region I as shown in Figure 2.26 (a) at 380μs. From this point, bolt slips

further to close the bolt-hole clearance without the requirement of additional loading. At the end

of this phase (at 401μs), bolt-metal contact is attained at bay region I and III as shown in Figure

2.26 (b). From this point, the majority of the load transference is through the bolt. From the

numerical model, the total time period of slipping is calculated to be 51μs within the specimen.

This observation can also be interpreted form the strain history of the bolt as shown in Figure

2.27.

32
Figure 2.25 Different phases on slipping event in aluminum/aluminum bolted joint
preloaded at 500lbf subjected to high loading rates.

Figure 2.26 Representation of bolt-metal contact during the event of slipping (a) at region I
and (b) at region I and III.

33
Figure 2.27 Correlation of strain-time history and bolt slip on aluminum/aluminum bolted
joint preloaded at 500lbf from finite element predictions.

2.4.3.4 FE results for tension test

Numerical analysis is performed for pull-off loading in SHPB for the specimen series S2 and S3

as mention in Table 2-2. The main objective of this study is to determine the stress concentrated

areas by visualizing Von Mises stress distribution and evaluating failure modes for different

bolted joint configurations. Influence of e/d and w/d ratio in the failure mode of the bolted joints

is investigated. Bearing strength and failure modes of joint assembly for varying e/d ratios are

presented in Figure 2.28 and Figure 2.29 respectively.

From Figure 2.28 and Figure 2.29, it can be seen that when we decrease the edge distance, the

failure mode changes from bearing to mixed mode of failures provided the w/d ratio maintained

in asymptotic region. The transition region for this type of joint assembly lie around e/d = 3 for

high strain rate loadings. Shear out failure, which is often considered to be dangerous to the

structural joint is observed for only e/d=1.

34
Figure 2.28 Variation of bearing strength of aluminum/aluminum bolted joint under
tension loading for different e/d ratios from finite element analysis.

Figure 2.29 Failure mode predictions of aluminum/aluminum bolted joint under tension
loading for different e/d ratios from finite element analysis.

Bearing strength and failure modes of joint assembly for varying w/d ratios are presented in

Figure 2.30 and Figure 2.31 respectively. With reference to the aforementioned plots: as the

35
width of the joint decreases (by keeping e/d ratio in asymptotic region), the predominant failure

mode changes from bearing to net tension. This transition point noticeably occurs at w/d = 2.828.

Figure 2.30 Variation of bearing strength of aluminum/aluminum bolted joint under


tension loading for different w/d ratios from finite element analysis.

Figure 2.31 Failure mode predictions of aluminum/aluminum bolted joint under tension
loading for different w/d ratios from finite element analysis.

36
2.5 Limitation

The finite element model is limited for higher preloading conditions. Also, only few parameters

are investigated from the numerical model: strength of the joint, preloading effect, effect of

slipping and the failure modes. The other influential factors such as contact stresses between the

faying surfaces, secondary bending of the joint, etc., are not explored.

2.6 Conclusion

In this study, structural behavior of single-lap metal-metal bolted joint is investigated for high

strain rate loading conditions. Experiments are carried out for compression loading in SHPB.

Refined 3D finite element model is developed using ABAQUS tool with complete experimental

setup which successfully predicts the behavior of joints for varying preloading conditions.

Preloads are applied in static (standard) solver and successfully imported in explicit solver for

analyzing the impact behavior of the joint. The results are in good compliance for higher torque

tightened cases. The study demonstrates that the joint slip, loading rate and bearing strength of

the bolted joints are more influenced by bolt clamping force subjected to impact loadings. The

torque applied to the bolts causes the compression on the metal adherends which increases the

frictional force between the faying surfaces. For time-dependent loading condition, this friction

is directly related to the loading rate and joint slip.

The validated FE model is further employed to investigate the joint failure modes for tension

loading. Bearing strength and failure modes are presented for varying geometrical

configurations. It shows that the failure mode of the bolted joints is entirely relying on joint

geometry. From this study, it is strongly recommended not to use e/d ≤ 3 and w/d ≤ 2.828 for

bolted joints subjected to high strain rate loadings.

37
REFERENCES

38
REFERENCES

[1] E. Godwin and F. Mathews, "A review of the strength of joints in fibre–reinforced," Compos
A Appl Sci Manuf, p. 11(3):155–60, 1980.

[2] G. Kretsis and F. Matthews, "The strength of bolted joints in glass fibre/epoxy laminates,"
Compos A Appl Sci Manuf, p. 16(2):92–102, 1985.

[3] P. Smith, K. Pascoe, C. Polak and D. Stroud, "The behavior of single-lap bolted joints in,"
Compos Struct, p. 6(1–3):41–55, 1986.

[4] A. VanderKlok, A. Dutta and A. Tekalur, "Metal to composite bolted joint behavior
evaluated at impact rates of loading," Composite Structures, vol. 106, p. 446–452, 2013.

[5] Y. Yan, W.-D. Wen, F.-K. Chang and et al, "Experimental study on clamping effects on the
tensile strength of composite plates with a bolt-filled hole," Compos: Part A, p. 1215–
1229, 1999.

[6] H.-S. Wang, C.-L. Hung and F.-K. Chang, "Bearing Failure of Bolted Composite Joints. Part
I: Experimental Characterization," Journal of Composite Materials, 1996, p.1284-1313.
[7] H. Kolsky, "An Investigation of the Mechanical Properties of Materials at very High Rates of
Loading," Proc. Phys. Soc, pp. 676-700, 1949.

[8] C. T. McCarthy, M. A. McCarthy, W. F. Stanley and V. P. Lawlor, "Experiences with


Modeling Friction in Composite Bolted Joints," Journal of Composite Materials, vol. 39,
p. 1881–1908, 2005.

[9] Zhou S, Wang Z, Wu X and Zhou J, "Experimental and numerical investigation on bolted
composite joint made by vacuum assisted resin injection," Compos Part B: Eng, p.
45(1):1620–8, 2013.

39
Chapter 3 Evaluating Hybrid Joint Behavior at Intermediate Strain Rates

3.1 Abstract

The importance of characterizing light-weight composite materials increases with increase in

implementation of composites in vehicle applications. In this study, the structural responses of

metal/composite joints (dissimilar material joints) under low and intermediate loading rates are

investigated in detail. Low velocity impact experiments are carried out on single-lap, single-bolt

Aluminum/S2 glass-SC 15 composite joints subjected to tension loading in instrumented Drop

Weight Impact Testing Machine (DWITM). Drop weight impact testing system is fundamentally

configured to simulate low velocity compressions which are comparable to automobile crashes.

However, tensile loading is induced in DWITM by a special kind of fixture designed and

developed to convert the loading from compression to tension. The effect of geometrical

parameters on bolted assembly is investigated for varying edge-to-bolt diameter ratio (e/d). The

study is further extended to evaluate the strength and failure modes of hybrid (bolted and

bonded) and adhesive joints. Experimental observation shows that the strength of the hybrid joint

is much higher than its bolted counterpart which is dominated by the load carrying capacity of

the bonded member. Also, the effect of hole-clearances, influence of adherend members and

environmental conditions on hybrid joint behaviors are analyzed in detailed.

3.2 Introduction

The use of advanced fiber reinforced composite materials for engineering structures in

aerospace, marine and automobiles has been increasing in recent years because of its excellent

mechanical properties and lighter weight. In structural assembly, joining composite materials to

other components are frequent and are often considered to be critical. A detail understanding of

40
joining techniques is therefore essential in composite assembly to prevent the catastrophic failure

of the structure during services.

3.2.1 Joining methods

Traditionally, joining two or more structural elements are achieved by mechanical fastening or

adhesively bonding technique that has its own advantages and disadvantages. Mechanical

fastening (uses bolts or rivets) method is considered to be the most ideal choice for joining

composite members due to its relative ease of installation and disassembly for inspection.

However, drilling holes in laminates may potentially damage the orientation of fibers in

composite configuration which requires specific attention [1], [2]. In addition to that, composite

materials are considerably brittle when compared to metals and the development of higher stress

regions at the bolt-hole surroundings significantly reduces the overall strength of the structure

[3]. On the other hand, adhesively bonded joints are proved to have exceptional static and fatigue

properties due to the absence of bolt-hole regions [4]. In this type of joining method, continuous

load transference can be achievable along the overlap length and hence considered to be the most

efficient joining method when compared to bolting [4]. On the contrary, it is highly impossible to

install adhesively bonded joints in critical structures due to the difficulty in detecting the onset of

failure within the adhesive layer. Also, the strength of the adhesively bonded joint is hugely

affected by the degradation of adhesive surface [5] within the bonding area.

Therefore, hybrid joints are developed in order to overcome the disadvantages of the

aforementioned joining techniques [6]. In hybrid joints, combination of mechanical bolts and

adhesives are employed together to reinforce the load carrying capability of the structural

assembly. In most practical applications, hybrid joints are used as a fail-safe technique, where if

one joining member fails, load can be taken over by the other member [7]. Also, it has been

41
shown that the inclusion of bolt in hybrid joint helps in enabling additional load transference by

the damaged adhesive bond by keeping any rupture to progress further [6], [7]. Thereby, the

catastrophic failure of the structure is prevented and enhances the long term performances of the

joint [7].

3.2.2 Testing methods

Understanding the behavior of joints under different loading rates is essential to optimize the

design and installation of joints in any structural members. Standard testing machines are

available to execute quasi-static analysis to predict the strength of the joints at very low strain

rates (1 s-1). To investigate the behavior of joints at higher strain rates (10 2 s-1 to 104 s-1), Split

Hopkinson Pressure Bar (SHPB) is generally used for both compression and tension. However, it

is extremely difficult to study the response of joints in intermediate strain rate loadings (10 0 s-1 to

102 s-1). A few experimental investigations have been carried out in drop tower testing machine

to characterize materials for intermediate rates of loading under compression [8], [9]. However,

performing tension test on dissimilar joints using drop tower is very challenging.

Previous works on composite joints are mainly focused on strength predictions on bolted

assembly subjected to static loading conditions. Smith et al. [10] observed the effects of the

geometrical parameters such as width and edge distance on ultimate strength and failure

mechanisms. Experiments are carried out by the authors to predict the strengths and failure

modes of single-lap bolted joints and compared with double lap joint configuration using CFRP

laminates. Based on theoretical approaches, Hart-Smith [3] predicted the substantial benefits of

using hybrid joints on improperly bonded structures. Kelly [11] carried out experimental and

numerical investigation with carbon-fiber reinforced plastic adherends to compare the

mechanical response of bolted, bonded and hybrid joints under static loading conditions. Kelly

42
[12] also studied the distribution of load in hybrid joints and investigated the effect of adhesive

stiffness on load sharing capability of the bolt.

3.2.3 Failure modes on composite joints observed from static testing

Figure 3.1 Typical failure modes observed in single-lap, single-bolt composite joints loaded
at the hole: (a) shear-out (b) cleavage (c) net tension and (d) bearing (Reproduced from
Figure 2.1).

Failure modes are experimentally investigated by number of authors under static loading

conditions [10], [13]. As demonstrated in Figure 3.1, typical failure modes observed in single-

lap, single-bolt composite joints under uniaxial tension load are classified as four types: shear-

out, cleavage, net tension and bearing failure [10],[14]. All failure modes should be taken into

consideration in designing structural joints. Shear-out, cleavage and net-tension failure modes

are catastrophic and detrimental that has to be avoided in engineering structures. However,

bearing failure is the preferred mode in mechanically fastened composite joints as it can

withstand loads beyond the onset of failure.

43
Few research works are published on bolted joints subjected to high loading rates [15]. However,

there is not much information available for joints under intermediate loading rate conditions. In

this study, the main objective is to predict the strength and failure modes of hybrid joint

configurations under low velocity tensile loading, with the focus on optimizing joint design for

vehicle applications. Impact testing is performed at room temperature in standard DWITM to

predict the strength of the joint and the results are compared against its static responses. Also, a

series of bolted and hybrid joints are tested at 60◦C to predict the strength degradation at elevated

temperature.

3.3 Experimental work

3.3.1 Static Testing Method

Static tests are conducted at room temperature using MTS 810 Material Test System. The test

specimens are tabbed in order to maintain axial alignment and clamped with 5 MPa gripping

pressure. Tensile force is applied to the bottom of the specimen with the displacement control

rate of 1 mm/min. To observe the failure mechanism, each joint specimen is loaded to its

ultimate failure load and the load-displacement responses are recorded. Videos and images are

captured using phantom camera which is used to examine the failure process of tested

specimens.

3.3.2 Low velocity Impact Testing Method

Drop weight tension testing is performed using Instron Dynatup 9250HV. This conventional

impact testing machine is mainly used to generate intermediate strain rate loadings in lab

environment which are comparable to actual automobile crashes. Impact on the sample can be

achieved by raising the mass of the system to a specific height and dropping it vertically on to

the testing specimen.

44
Figure 3.2 Drop tower test setup (left), fixture setup with holding specimen (middle) and
joint specimen (right).

Along with variable impact height, the tower is also capable of varying mass, velocity and

impact energy. The movement of the suspended mass is channeled by the columns that are held

in position as shown in Figure 3.2. As the name signifies, DWITM are originally configured to

generate compression loading. In the current study, this instrumented loading device is used to

induce tension load by manufacturing a special type of fixture which can capable of changing the

load from compression to tension. The schematic representation of the tensile fixture is shown in

Figure 3.3.

In the fixture, the joint assembly is clamped with the top of the specimen attached to the

stationary base and the bottom of the specimen attached to the sliding platform which takes the

impact, thereby inducing tensile loading on the tested samples. Soft materials, so called pulse

shapers, are used on top of the tensile fixture in order to achieve force equilibrium across the

specimen length. Load sensors are placed at the top and bottom of the specimen which are used

45
to record the force history of the sample. High speed images are taken at the rate of 15000

frames per second using Phantom V12 high speed camera to examine joint responses. Tests at

elevated temperatures are conducted in environmental conditioning unit equipped with Dynatup

9250HV. The specimens are heated to 60◦C in the heat chamber for 15 min prior to testing to

ensure thermal equilibrium.

Figure 3.3 Fixture used to convert compression to tension load in drop tower machine.

46
3.3.3 Specimen configurations and materials

The schematic representation of the tested specimen and joint parameters are shown in the

Figure 3.4. Single-lap, metal-composite joint configurations are tested to investigate the

mechanical response of structural joints under static and intermediate loading rates. In all the

tested specimens, aluminum of thickness 0.125 in. and e/d ratio of 2 is used as the metal

adherend (Figure 3.5). To study the influence of geometrical parameters, S2 glass/SC 15 epoxy

composite laminates of thickness 0.145 in. are used as the composite adherend for varying e/d

ratios from 1.0 to 4.0. The material properties of the composite laminate are shown in Table 3-1.

The efficiency of three different joining methods (bolted, bonded and hybrid) are examined using

aluminum/S2 glass-SC 15 composite joints. For all bolted and hybrid specimens, clamping force

of 40 in.lb is applied to the bolt of diameter 0.25 in. FM-94K film adhesive is used for bonded

and hybrid joints with the bond area of 0.5 in. x 1 in. The material property of the adhesive

member is provided by the manufacturer. Specimens are heated at 3°C/min and cured at 127°C

for 2 hrs. Special fixtures are employed to preserve the alignment of joint and to maintain the

clamping pressure on the joint assembly (Figure 3.6). The behavior of hybrid joints with e/d =1

is compared for three different hole-clearances: (a) tight fit (b) 0.006 in. and (b) 0.03 in. Also, the

effect of composite adherend is investigated by varying the composite material for bolted and

hybrid configurations with e/d = 1 and 4. E-glass plain weave ply with a matrix of black epoxy

resin of thickness 0.094 in. is used for comparative study. Investigation at elevated temperatures

is conducted on aluminum/S2 glass-SC 15 using bolted and hybrid joints with e/d = 1 and 3.

47
Figure 3.4 Hybrid joint specimen configuration used for impact analysis and geometrical
parameters: Front view (shown in left) and side view (shown in right).

Figure 3.5 Bolted joint geometry used for impact analysis.

48
Table 3-1 Material properties of S2 glass-SC 15 epoxy laminates [16].

Figure 3.6 Aluminum/S2 glass-SC 15 hybrid samples placed on the specially designed
fixture to maintain alignment and constant pressure.

3.4 Experimental Results and Discussion

3.4.1 Experimental observations from static analysis

3.4.1.1 Response of bolted joints

The load-displacement curve of the bolted specimen with edge distance 1d is shown in Figure

3.7. Static response of the bolted joint can be classified as four stages: initial loading, slip, post

slip and bearing failure [17]. In stage 1 (AB), linear behavior of the load-displacement curve is

observed which represents the slip resistance caused by the contact force between the adherend

members. Frictional force between the matting surfaces varies with the bolt preload [18].

49
Figure 3.7 Load-displacement curve of Aluminum/S2 glass-SC 15 bolted joint with edge
distance 1d under static loading.

In stage 2 (BC), a constant steady load is observed as the displacement increases. This stage

indicates the initiation of bolt slipping from exceeding frictional force. The length of slipping

entirely depends on the hole clearance and bolt clamping force. In stage 3(CD), non-linear

response of the load is observed which represents the post slip behavior of the joint. Within this

region, the contact between bolt and specimen is achieved that initiates and develops bearing

failure on further loading. The final stage (DE) represents the sudden reduction in load after

exceeding the maximum capacity of the joint.

3.4.1.2 Effect of e/d ratio

To investigate the effect of edge distance in bolted joint configuration, a total of 16 specimens

are tested with varying edge distance from 1d to 4d. Figure 3.8 illustrates the relationship

between the failure force and e/d ratio of the bolted configuration. The error bars indicate 95%

confidence for 4 samples. Results show that the load carrying capacity of the bolted joints is

directly proportional to the edge distance. Also, it is observed that the specimens with higher e/d

50
ratio, requires more time to reach the peak force which is attributed to the material overlap

length. Failure force of the bolted configuration with edge distance 4d is 172% greater than that

of 1d.

Figure 3.8 Failure loads of Aluminum/S2 glass-SC 15 bolted joints with different e/d ratios
under static loading.

Also, with reference to Figure 3.9, it can be seen that the mode of failure changes from shear

out to bearing as the edge distance increases. Based on the observations it is evident that the

change in failure mode has a clear dependency on the edge distance and there exists a transition

region where the bearing mode of failure starts to dominate. The experiments predict that the

transition region for bolted joint configuration lies around e/d = 3.

Many research works has been carried out to investigate the edge effect on composite joints [10],

[15], [19], and majority of authors seem to agree that a certain minimum edge distance

(transition region) is required to reach the dominating mode of failure. However, in designing

structural joints, the transition e/d ratio has to be decided based on additional geometrical

parameters such as width, thickness, and lay-up sequence [14] that needs to be explored in

future.

51
Figure 3.9 Failure modes of Aluminum/S2 glass-SC 15 bolted joints with different e/d ratios
under static loading.

3.4.1.3 Response of bonded joints

Load-displacement curve for bonded joint with bonding area of 0.5 in. x 1 in. is shown in Figure

3.10. Static response of the bonded joint can be classified as two stages. Initial stage (A”B”)

represents the slip resistance of the joint up to the maximum load carrying capacity of the

adhesive member. In bonded joint, slip resistance is increased by the presence of adhesive

member which helps in smooth load transference when compared to bolted configuration.

However, after exceeding its maximum load carrying capacity, bonded joint fails

catastrophically (B”C”). With reference to Figure 3.10, the strength of the bonded joint is found

to be 7.52 kN and it might depend on the stiffness of the adhesive [11]. Further investigation

needs to be done to verify the load carrying capacity of the joint using adhesives of varying

stiffness.

52
Figure 3.10 Load-displacement curve of Aluminum/S2 glass-SC 15 bonded joint with bond
area 0.5 in. x 1 in. under static loading (load bottom to top).

3.4.1.4 Failure modes on bonded joints

The failure modes on bonded joints can be classified as four types corresponding to the

definition shown in Figure 3.11 [20]. Aluminum/S2 glass-SC 15 bonded joints evaluated for

static loading shows (Figure 3.12) a mixed mode of failure which indicates partial adhesive

failure on both adherend members.

Figure 3.11 Typical failure modes definition of an adhesive bonded joint: (a) cohesive
failure, (b) adhesive failure, (c) mixed failure and
(d) adherend failure (Reproduced from [20]).

53
Figure 3.12 Failure modes of Aluminum/S2 glass-SC 15 bonded joints with bond area 0.5
in. x 1 in. under static loading.

3.4.1.5 Response of hybrid joints

Figure 3.13 Load-displacement curve of Aluminum/S2 glass-SC 15 hybrid joint with edge
distance 1d under static loading (load bottom to top).

The load-displacement curve for hybrid joint with edge distance 1d is shown in Figure 3.13.

Similar to the bolted configuration, behavior of hybrid joints are classified as 4 stages [17]. The

first stage (A’B’) indicates the slip resistance of the joint up to the failure of adhesive member.

In this stage, the complete load transference is established by the bonded member without the

54
contribution of frictional force between the contact surfaces. In the second stage (B’C’), sudden

failure of adhesive member is observed followed by the joint slip. The slipping of bolt is not

much evident in hybrid joints. This might be due to the coupled behavior of joint slip and

adhesive failure. In the third stage (C’D’), non-linear behavior of the load is observed up to the

maximum load carrying capacity of the bolt along with the gradual development of bearing

failure. The final stage (DE) represents the sudden failure of the joint specimen.

3.4.1.6 Effect of e/d ratio

Figure 3.14 Failure load predictions of the Aluminum/S2 glass-SC 15 hybrid joints with
different e/d ratios under static loading.

Total of 16 specimens are tested with varying edge distance from 1d to 4d to examine the effect

of edge distance. Figure 3.14 shows the experimental failure loads of the hybrid joints with

different e/d ratios. As can be seen, maximum load carrying capacity of the hybrid joints are

nearly identical for varying edge distances. This is because; the strength of the hybrid joints

depends entirely on adhesive stiffness.

Also, with reference to Figure 3.15, it can be seen that the failure pattern changes from shear

out to bearing as the edge distance increases and reaches a transition region at e/d = 3. Along

55
with the typical failure modes (Figure 3.1), failure of adhesive member contributes mixed mode

of failure (Figure 3.11) in all hybrid configurations as shown in Figure 3.12.

Figure 3.15 Failure modes of Aluminum/S2 glass-SC 15 hybrid joints with different e/d
ratios under static loading.

Figure 3.16 Failure modes of Aluminum/S2 glass-SC 15 hybrid joints with e/d =1 and bond
area of 0.5 in. x 1 in. under static loading.

3.4.1.7 Effect of joint type

Figure 3.17 illustrates the summary of joint failure loads for bolted, bonded and hybrid

configurations. It is evident that the load carrying capacity of the hybrid joints is much higher

than that of the bolted joints for all edge distances. Failure load of hybrid joints with e/d =1 is

56
nearly 3 times higher than that of the bolted joints with similar configuration and is 1.15 times

higher than that of bolted joints with e/d = 4.

Figure 3.17 Failure load comparison of Aluminum/S2 glass-SC 15 bolted, bonded & hybrid
joints under static loading

Also, it can be seen that the failure loads of the hybrid joints are comparable to the ultimate

failure loads of the adhesive joints. However, hybrid joints are considered to be superior to

bonded joints with respect to failure pattern. This can be interpreted from the load-displacement

curves of the aforementioned joints. As can be seen from Figure 3.18, the load distribution in

hybrid joint is typically similar to that of bonded joint before the onset of adhesive failure. Once

the bonded member fails, the behavior of hybrid joint is similar to that of the bolted joint and the

load transference is achieved by the bolted member alone. Joint load is sustained by the bolt

which results in gradual and progressive failure. On the other hand, failure of bonded joints is

catastrophic which has a detrimental effect on any structure and hence needs to be avoided. This

57
sort of hybrid joint behavior is in disagreement with the observations by Kweon et al [20] where

the bearing mode of failure occurs before the adhesive member fails.

Figure 3.18 Force vs. displacement plot of Aluminum/S2 glass-SC 15 joints under static
loading (a) bolted joint with e/d = 1, (b) bonded joint with bond area 0.5 in. x 1 in. and (c)
hybrid joint with e/d =1.

3.4.2 Experimental observations from impact analysis

3.4.2.1 Force equilibrium

During the event of impact loading, maintaining constant force or at least nearly constant force

throughout the specimen is always a challenging task. Failure to attain force equilibrium may

result in an instantaneous and premature damage in the specimen which may invalidate the test

results. Achieving uniform force distribution within the specimen is therefore essential in order

58
to attain gradual and progressive failure. In the current experiment, the direct impact is made on

top surface of the tensile fixture as shown in Figure 3.3. The resulted impact pulse causes the

movable frame to slide down and induces force at the bottom of the specimen. The induced

tensile force will then propagate along the length of the specimen that requires a short period of

time to achieve uniform distribution of force within the specimen.

Table 3-2 Dimensions of pulse shaping materials used to achieve force equilibrium within
the specimen under impact loading.

In realistic conditions, premature fracture on the specimens is observed before it reaches force

equilibrium. This can be eliminated by adding a softer material, called pulse shapers [14]

between the impact tup and the top surface of the fixture. Determining the suitable type of pulse

shaping material is purely a trial and error method.

In this study, force equilibrium is verified for a myriad of different pulse shaping materials but

the details are presented only for four cases. Investigated pulse shaping materials are categorized

from hard to soft rubber material and are listed in Table 3-2. Conventional dog-bone aluminum

sample of thickness 0.125 in. is used to examine the force equilibrium under tensile loading

conditions. Force data acquired by the top and bottom load cells are presented in Figure 3.19 for

four pulse shaping materials. The differences between the peak force from the top and bottom

load cell measurements are significantly large for hard rubber material as shown in Figure 3.19

59
(a). Continuous development in force equilibrium can be seen from Figure 3.19 (b) and (c)

when medium rubber materials are used. On using soft rubber material, about 80% of force

transmission is achieved which demonstrates the uniform distribution (approximately) of force

throughout the length of the specimen (Figure 3.19 (d)). Therefore Latex rubber of thickness

0.25 in. is used for later analysis.

(a) (b)

(c) (d)

Figure 3.19 Force histories of dog-bone aluminum specimen from top and bottom load cell
on using different pulse shaping materials: (a) Buna-N-rubber, (b) polyurethane, (c)
polypropylene and (d) latex rubber.

60
3.4.2.2 Response of bolted joints

Impact behavior of bolted joint configuration is illustrated in Figure 3.20. Similar to static

observation, all four phases on joint behavior under intermediate loading rates are evident with a

substantial bolt rotation of about 5° is observed along with bolt slipping as shown in Figure

3.21. The reason could possibly due to the significant bending on the single-lap joints under

impact rates of loading. This issue can be eliminated by using double-lap configurations [1], but

it is virtually impossible to replace every single-lap joints in practical application.

Figure 3.20 Force history of Aluminum/S2 glass-SC 15 bolted joint with edge distance 1d
under impact loading from top load cell (load right to left).

3.4.2.3 Effect of e/d ratio

A total of 12 specimens are tested for varying edges distances from 1d to 4d and the

experimental failure loads are shown in Figure 3.22. The error bars indicate 95% confidence for

3 samples. Results show that the maximum load carrying capacity of the bolted joint increases

61
with edge distance and plateaus out at higher e/d ratios. The failure load of bolted joint with e/d

= 3 is at least twice as high as the joint failure load of e/d = 1. Furthermore, it can be observed

that the load transfer curve tends to elongate with the increasing overlap area as more time is

required to reach complete failure.

Figure 3.21 Bolt transition & rotation of Aluminum/S2 glass-SC 15 bolted joint with edge
distance 1d under impact loading (load right to left).

Also, with reference to Figure 3.23, it is evident that the bearing failure mode starts to occur for

higher e/d dimensions. A significant delamination on the composite materials is observed when

the loading rate is inflicted. This can be attributed to the rate sensitive properties of S2 glass-SC

15 composite materials at which they are loaded. The potential reason could also be due to the

stacking sequence of the laminate [21] that affects the interlaminar shear stresses around the hole

which demands further investigation.

62
Figure 3.22 Failure loads of Aluminum/S2 glass-SC 15 bolted joints with different e/d
ratios under intermediate loading rates.

Figure 3.23 Failure modes of Aluminum/S2 glass-SC 15 bolted joints with different e/d
ratios under impact loading.

63
3.4.2.4 Response of bonded joints

Impact response of the bonded joint with bonding area of 0.5 in. x 1 in. is shown in Figure 3.24.

Unlike static behavior, crack initiation on adherend member is observed prior to the peak load in

A”B” region. The crack initiation in the adherend member can be attributed to the bending of the

joint as illustrated in Figure 3.25. This crack develops further which causes complete adherend

failure in a simply bonded joint. The peak load (B1”) represents the failure of adherend member

and found to be 5.60kN.

Figure 3.24 Force history of Aluminum/S2 glass-SC 15 hybrid joint with edge distance 1d
under impact loading from top load cell (load right to left).

3.4.2.5 Failure modes on bonded joints

Adhesively bonded joints are tested to observe the failure modes under intermediate loading

rates. Among four, three samples exhibit adherend failure (Figure 3.26) which are used to

calculate average joint failure load. However, one specimen failed due to adhesive failure which

is neglected in data analysis. The potential reason for adhesive failure could be due to inefficient

surface preparation.

64
Figure 3.25 Schematic of joint slip (yellow line) and bending (blue line) of Aluminum/S2
glass-SC 15 bonded joint with bond area 0.5 in. x 1 in. under impact loading (load right to
left).

Figure 3.26 Failure modes of Aluminum/S2 glass-SC 15 bonded joints with bond area 0.5in.
x 1in. under impact loading.

65
Contact between the adhesive and adherend members are mainly defined by the surface

treatement. If the process is not done appropriately, adhesion failure could posibly occur which

significantly reduces the load carrying capacity of the joints. In this investigation, all hybrid and

bonded joints are prepared by abrading the adherend surfaces using fine grit sand paper prior to

bonding. To observe the influence of surface treatment, further investigation needs to be done in

future.

3.4.2.6 Response of hybrid joints

The behavior of hybrid joints is classified as four stages (Figure 3.27). In the initial stage

(A1’B1’), slip resistance of the joint is offered by the combination of bolt and adherend member.

This can be explained by comparing its failure pattern with bonded configuration (Figure 3.24).

Unlike bonded joints, there is no delamination in the composite laminates during initial loading.

Also, the presence of bolt prevents the bending on adherend member.

Figure 3.27 Force history of Aluminum/S2 glass-SC 15 hybrid joint with e/d = 1 under
impact loading (load right to left).

66
Partial debonding on adhesive constituent is spotted at first peak (B1’) which is in contrast to the

static observation where complete failure of adhesive member is observed. The second stage

(B1’C1’) indicates the gradual slip on the joint into the bearing damage. The structural integrity

of the joint sustains partially by adhesive and by bolt member in C1’D1’ region followed by the

complete failure of adhesive at D1’. Once adhesive fails, the joint load is sustained by the bolt

member till it reaches its ultimate failure load.

3.4.2.7 Effect of e/d ratio

Figure 3.28 shows the experimental failure loads of the hybrid joints are almost constant for

varying edge distances. Also, with reference to Figure 3.29, it can be seen that the failure mode

changes from shear out to bearing as the edge distance increases with significant inter-laminar

and mixed mode of failure (Figure 3.11). The failure mode ultimately reaches a transition

region at e/d = 3, where the bearing mode of failure starts to dominate.

Figure 3.28 Failure load predictions of the Aluminum/S2 glass-SC 15 hybrid joints with
different e/d ratios under impact loading.

67
Figure 3.29 Failure modes of Aluminum/S2 glass-SC 15 hybrid joints with different e/d
ratios under static loading.

3.4.2.8 Effect of joint type

Figure 3.30 illustrates the summary of joint failure loads of bolted, bonded and hybrid

configurations under intermediate rates of loading. It can be seen that the failure loads of hybrid

joints are significantly higher for lower e/d ratios when compared to bolted joints of similar

dimensions. Howerver, for higher e/d ratios, there is no significant differences in strength is

observed. Also, it can be seen that the load carrying capacity of the hybrid joints with lower e/d

ratios are comparable to bolted joint loads with higher e/d ratios. Failure patterns of hybrid joints

are similar to that of bolted configuration for all e/d ratios except for the increase in damage area

in where bearing mode of failure dominates. This is due to the combined effect of loading rate

and stiffness of the adhesive member. Adhesively bonded joints fail predominantly by adherent

failure. Hence, it will not be logical to compare the ultimate joint failure load of bonded with

bolted or hybrid configurations.

68
Figure 3.30 Failure load comparison of Aluminum/S2 glass-SC 15 bolted, bonded & hybrid
joints under impact loading.

3.4.2.9 Effect of loading rates

Figure 3.31 illustrates the summary of joint failure loads for bolted, bonded and hybrid

configurations under static, intermediate and high loading rates [22]. The increasing trend in

joint failure force with increasing loading rate indicates that the material properties are a function

of loading rates which are in good agreement with the results by Daimaruya et al [23]. For all

loading conditions, hybrid joints exhibit larger load carrying capacity. Under static and

intermediate loading rates, hybrid joint with e/d = 1 has a comparable strength to bolted joint of

higher e/d dimensions. Also, it can be seen that the hybrid joints with e/d = 1 is the least rate

sensitive configuration.

69
Figure 3.31 Failure load comparison of Aluminum/S2 glass-SC 15 bolted, bonded & hybrid
joints under static, intermediate and high loading rates.

3.4.2.10 Effect of hole-clearance

The influence of hole-clearance in hybrid joints is investigated for three cases: tight fit, 0.006 in.

and 0.03 in. As demonstrated in Figure 3.32, increasing the hole clearance does not play a

dominant role on joint failure load. However, the magnitude of bolt slip increases with increase

in bolt-hole clearance. For an efficient joint design, it is essential to define an optimum range for

bolt-hole clearance under impact loading condition. From the experimental observations, 0.03 in.

clearance is believed to be efficient since tight fit is virtually impossible to achieve in practical

scenarios.

70
Figure 3.32 Behavior of Aluminum/S2 glass-SC 15 hybrid joints with e/d = 1 for varying
hole clearances under intermediate loading rates.

3.4.2.11 Effect of environmental conditions

Experiments are conducted for bolted and hybrid joints with e/d = 1 and 3 at 60°C to investigate

the joint sensitivity to environmental temperature. A total of 32 samples are tested and the results

are illustrated in the Figure 3.33. Reduction in load carrying capacity of joints at elevated

temperature is observed for lower e/d ratios. However, the joint failure load is insensitive to

environmental conditions for larger e/d values under intermediate loading rates. Joints with e/d =

1 exhibits shear out failure mode which initiates from exceeding the compressive strength of the

fiber and matrix along the loading direction. On the other hand, failure in higher e/d ratios is not

dominated by matrix properties. Therefore, it is believed that the reduction in ultimate load for

lower edge distances with increase in temperature can be attributed to the degradation of the

matrix strength in composite laminates at elevated temperatures. The operative temperature is in

the range of glass transition temperature (Tg, 95°C [23]) of SC 15 epoxy which causes gradual

softening of matrix constituent that affects the load carrying capacity of the joints. However,

71
glass transition temperature of S2 glass (Tg, 1056°C [24]) is way beyond the operating range

which is not attributed to the deduction in strength. A fairly small difference in hybrid joint

failure load is observed for higher e/d ratios at 60°C, which may due to the softening of the

adhesive with increase in temperature.

Figure 3.33 Failure load comparison of Aluminum/S2 glass-SC 15 bolted & hybrid joints
with e/d = 1 and 3 between room temperature and 60°C.

3.4.2.12 Effect of composite adherend

To investigate the effect of joining substrates, experiments are conducted on single-lap,

aluminum/E-glass ply epoxy resin composites for bolted and hybrid configurations with e/d = 1

and 4. Four joint specimens are tested for each model to determine the average joint failure force.

The bearing strength of the joints is caluclated as the joint failure load divided by the thickness

of the composite laminates times the diameter of the bolt-hole, regardless of the joining method.

This method of comparison is carried out purely for convenient purposes.

72
Figure 3.34 Bearing strength comparison between Aluminum/S2 glass-SC 15 and
Aluminum/E-glass bolted joints with e/d =1 and 4 under intermediate loading rates.

Figure 3.35 Bearing strength comparison between Aluminum/S2 glass-SC 15 and


Aluminum/E-glass hybrid joints with e/d =1 and 4 under intermediate loading rates.

With reference to Figure 3.34, 35% increase in the bearing strength of the aluminum/E-glass

bolted joint with e/d = 4 is observed. However, there is no considerable difference in bearing

strength for lower e/d ratios. This can be explained based on the failure patterns on the

corresponding adherends Figure 3.36. From experiments, it is observed that S2-glass-SC 15

composite exhibits significant delamination for higher e/d ratios under impact loading rates.

However, in E-glass specimens no substantial delamination can be seen. Observation of increase

73
in bearing strength in E-glass laminates might be attributed to the increase in interlaminar shear

strength of the material when compared to S2-glass-SC 15 composites. For the same reason,

aluminum/E-glass hybrid joints exhibits higher bearing strength when compared to

aluminum/S2-glass-SC 15 composite of similar joint configurations as shown in Figure 3.35.

Figure 3.36 Failure mode comparison between Aluminum/S2 glass-SC 15 and


Aluminum/E-glass bolted and hybrid with e/d = 1 and 4 under impact loading rates.

3.5 Conclusion

Quasi-static and low velocity impact tests are conducted on aluminum/s2glass-SC15 single lap

joints. The strength of the joints is evaluated and the failure modes are characterized for three

types of joint configurations: mechanical fastening, bonding, and hybrid joints. It is shown that

the strength of the hybrid joints with shorter edge distances is stronger than the bolted

configuration with large edge distances. Hybrid joint shows improved load carrying capability by

the presence of adhesive member. And on post-adhesive failure; joint load is sustained by the

bolt thereby prevents catastrophic failure of the joint. Modes of failure for both bolted and hybrid

joints show asymptotic behavior at e/d =3 where bearing mode of failure dominates. Significant

74
delamination on joints under impact rates of loading is observed which is attributed to the rate

sensitivity of the material. At elevated temperatures (60°C), there is a 25-35% reduction of

strength for joints with smaller edge distances while joint strength for large edge distances is

invariant with respect to environmental conditions attributed to the gradual softening of matrix

constituent in composite laminates. The effects of hole clearance, and adherend material are also

identified as critical parameters which should be considered on designing of joints that may

undergo impact rates of loading.

75
REFERENCES

76
REFERENCES

[1] A. . Mouritz, E. Gellert, P. Burchill, and K. Challis, “Review of advanced composite


structures for naval ships and submarines,” Compos. Struct., vol. 53, no. 1, pp. 21–
42, Jul. 2001.

[2] P. K. Mallick, Composites Engineering Handbook. CRC Press, 1997.

[3] L. J. Hart-Smith, “Design Methodology for Bonded-Bolted Composite Joints. Volume


I. Analysis Derivations and Illustrative Solutions,” Feb. 1982.

[4] W. R. Broughton, L. E. Crocker, and M. R. L. Gower, Design Requirements for Bonded


and Bolted Composite Structures. National Physical Laboratory, 2002.

[5] L. J. HART-SMITH, “Bonded-bolted composite joints,” J. Aircr., vol. 22, no. 11, pp.
993–1000, 1985.

[6] L. J. Hart-Smith, “An Engineer’S Viewpoint on Design and Analysis of Aircraft


Structural Joints,” Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng., vol. 209, no. 2, pp.
105–129, Apr. 1995.

[7] D. M. Jan R. Weitzenböck, “Science and Technology of Bolt-Adhesive Joints,” pp.


177–199, 2011.

[8] P. O. Sjoblom, J. T. Hartness, and T. M. Cordell, “On Low-Velocity Impact Testing of


Composite Materials,” J. Compos. Mater., vol. 22, no. 1, pp. 30–52, Jan. 1988.

[9] J. D. Winkel and D. F. Adams, “Instrumented drop weight impact testing of cross-ply
and fabric composites,” Composites, vol. 16, no. 4, pp. 268–278, Oct. 1985.

[10] P. A. Smith, K. J. Pascoe, C. Polak, and D. O. Stroud, “The behaviour of single-lap


bolted joints in CFRP laminates,” Compos. Struct., vol. 6, no. 1–3, pp. 41–55, 1986.

[11] G. Kelly, “Quasi-static strength and fatigue life of hybrid (bonded/bolted) composite
single-lap joints,” Compos. Struct., vol. 72, no. 1, pp. 119–129, Jan. 2006.

77
[12] G. Kelly, “Load transfer in hybrid (bonded/bolted) composite single-lap joints,”
Compos. Struct., vol. 69, no. 1, pp. 35–43, Jun. 2005.

[13] C. Cooper and G. J. Turvey, “Effects of joint geometry and bolt torque on the structural
performance of single bolt tension joints in pultruded GRP sheet material,” Compos.
Struct., vol. 32, no. 1–4, pp. 217–226, 1995.

[14] G. Kretsis and F. L. Matthews, “The strength of bolted joints in glass fibre/epoxy
laminates,” Composites, vol. 16, no. 2, pp. 92–102, Apr. 1985.

[15] A. VanderKlok, A. Dutta, and S. A. Tekalur, “Metal to composite bolted joint behavior
evaluated at impact rates of loading,” Compos. Struct., vol. 106, pp. 446–452, Dec.
2013.

[16] J. R. Xiao, B. A. Gama, and J. W. Gillespie Jr., “Progressive damage and delamination
in plain weave S-2 glass/SC-15 composites under quasi-static punch-shear loading,”
Compos. Struct., vol. 78, no. 2, pp. 182–196, 2007.

[17] N. D. Hai and H. Mutsuyoshi, “Structural behavior of double-lap joints of steel splice
plates bolted/bonded to pultruded hybrid CFRP/GFRP laminates,” Constr. Build.
Mater., vol. 30, pp. 347–359, May 2012.

[18] H.-S. Wang, C.-L. Hung, and F.-K. Chang, “Bearing Failure of Bolted Composite
Joints. Part I: Experimental Characterization,” J. Compos. Mater., vol. 30, no. 12, pp.
1284–1313, Aug. 1996.

[19] E. W. Godwin and F. L. Matthews, “A review of the strength of joints in fibre-


reinforced plastics: Part 1. Mechanically fastened joints,” Composites, vol. 11, no. 3,
pp. 155–160, Jul. 1980.

[20] J.-H. Kweon, J.-W. Jung, T.-H. Kim, J.-H. Choi, and D.-H. Kim, “Failure of carbon
composite-to-aluminum joints with combined mechanical fastening and adhesive
bonding,” Compos. Struct., vol. 75, no. 1–4, pp. 192–198, Sep. 2006.

[21] N. J. Pagano and R. B. Pipes, “The Influence of Stacking Sequence on Laminate


Strength,” J. Compos. Mater., vol. 5, no. 1, pp. 50–57, Jan. 1971.

78
[22] D. Gonzalez, “Mechanical behavior of metal-composite joints under impact rates of
loading,” Michigan State University, 2014.

[23] I. M. M. Michael L. Wang, “Materials Characterization of High-Temperature Epoxy


Resins: SC-79 and SC-15/SC-79 Blend,” p. 34, 2011.

[24]R. Alagirusamy and A. Das, Technical Textile Yarns. Elsevier, 2010.

79
Chapter 4 Conclusions and Recommendations

Joining of two structural members is commonly achieved by bolting, bonding, or a combination

of the two methods (hybrid) in engineering structures such as aircraft, marine and automotive. If

not designed properly, dynamic loading of these joints can be critical to the integrity of the

structure, as failure initiates at fastener-material interfaces. This demands a thorough

understanding of the structural response of joints subjected to various loading conditions.

4.1 High strain rate impact testing

In this study, enhanced three dimensional (3D) finite element (FE) models are developed for

single-lap, single-bolt, metal-metal joints in order to observe their structural behavior when

subjected to impact loading in a Split Hopkinson Pressure Bar (SHPB). FE solutions are

compared against experimental results and are found to be in good agreement for compression

loading in SHPB. The study demonstrates that the joint slip, loading rate and bearing strength of

the bolted joints are more influenced by bolt clamping force subjected to impact loadings. The

torque applied to the bolts causes the compression on the metal adherends which increases the

frictional force between the faying surfaces. For time-dependent loading condition, this friction

is directly related to the loading rate and joint slip. The validated FE model is further employed

to investigate the joint failure modes for tension loading. Bearing strength and failure modes are

presented for varying geometrical configurations. It shows that the failure mode of the bolted

joints is entirely relying on joint geometry. From this study, it is strongly recommended not to

use e/d ≤ 3 and w/d ≤ 2.828 for bolted joints subjected to high strain rate loadings.

4.2 Static and low velocity impact testing

Quasi-static and low velocity impact tests are conducted on aluminum/s2glass-SC15 single lap

joints. The strength of the joints is evaluated and the failure modes are characterized for three

80
types of joint configurations: mechanical fastening, bonding, and hybrid joints with various

geometric configurations. Intermediate loading rate impact testing is conducted on a Drop

Weight Impact Testing Machine (DWITM) where tensile loading is applied by a specially

designed fixture. Results show that at room temperature, the strength of the hybrid joints with

shorter edge distances is greater than or comparable to the strength of bolted joints with larger

edge distances. Modes of failure for both bolted and hybrid joints show asymptotic behavior at

e/d =3 where bearing mode of failure dominates. Significant delamination on joints under impact

loading can be attributed to material rate sensitivity. At elevated temperatures (60°C), there is a

25-35% reduction of strength for joints with smaller edge distances while joint strength for large

edge distances is invariant with respect to environmental conditions.

4.3 Recommendations

The developed finite element model can be further employed to understand other influential

parameters such as contact stresses between the faying surfaces and secondary bending of the

joint under impact rates of loading. Also, material constituent on finite element model can be

modified to study the complex behavior of dissimilar material joints.

81

You might also like