4 point bend testing

Upload: anand-ras

Post on 19-Feb-2018

212 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/23/2019 4 point bend testing

    1/6

    High-Speed Bend Test Method and Failure Prediction for Drop Impact Reliability

    Seah S.K.W., Wong E.H., Mai Y.W.*, Rajoo R., Lim C.T.**

    Institute of Microelectronics, Singapore

    *University of Sydney

    **National University of Singapore

    AbstractThe objective of this study is to obtain experimental

    failure models governing solder joint failure during drop

    impact testing of board assemblies. A high-speed bend tester

    was developed to perform displacement-controlled bend tests

    of board assemblies at the high flexing frequencies of drop

    impact. These test frequencies and amplitudes are not

    achievable by conventional universal testers. Experimental

    data was obtained for various PCB strain amplitudes, flexural

    frequencies, solder alloys and pad finishes. Results from the

    high-speed bend tests are used to construct constant-

    amplitude power law fatigue curves. Solder joint reliability is

    found to be dependent on the test frequency, and therefore

    strain rate. The experimental failure data from these high-speed bend tests are a required basis for a drop impact failure

    criterion which can take into account frequency and

    amplitude effects and which is general enough to be applied

    to product level testing.

    1. Introduction

    1.1 Board level testing

    Interest in the drop impact reliability of electronic

    packaging is motivated by 1) the widespread use of portable

    electronics; 2) miniaturization trends which can lead to

    smaller and more fragile interconnections; and 3) the use of

    lead-free solder, some of which have been shown to result in

    fragile joints under drop impact conditions [1-4]. For several

    years now, the standard in the electronic packaging industryfor board assembly drop impact testing has been the JESD22-

    B111 standard [5] by JEDEC. This is a board-level test in

    which a standard board assembly test vehicle is used, in

    contrast to the product-level test in which actual products

    such as mobile phones are tested. The board-level test is

    shown in Fig. 1.

    A significant number of experimental and numerical

    simulation studies have been performed based on the JEDEC

    board-level drop test [6-11]. Most of these studies involve: 1)

    comparison studies of the impact performance for different

    packaging parameters and different solder joint locations on

    the JEDEC board [6-8]; and 2) finite element (FE) modeling

    methodologies for simulating the JEDEC drop test [6-11].Failure predictions using the JEDEC drop test standard are

    typically based on the number of drops to failure at a certain

    reference load. An example of a reference load could be the

    initial peak value of the interconnection stress waveform

    obtained from drop impact modeling [6]. These failure

    predictions rely on the assumption that in a comparison study

    of drop performance, the PCBs of the test vehicles being

    compared have the exactly the same construction and

    boundary (mounting) conditions and therefore have the same

    dynamic behavior. Only if the dynamic responses of the PCBs

    are almost identical can the number of drops to failure be used

    as a comparative measure of reliability.

    Figure 1: Board-level drop test showing dynamic

    flexing of PCB

    Figure 2: FEA plot (magnified) of board assembly flexure

    showing deformation of outermost solder joint

    1.2 Failure criteria studies

    There have also been several studies and proposals aimed

    at developing failure criteria which are more fundamental and

    less dependent on the PCB dimensions and mounting. Such

    failure criteria can be general enough to be applicable to

    product level-testing rather than certain board-level test

    setups.

    Zhu and Williams [12] proposed a plastic strain criterion.

    In FE modeling, the plastic strain is a representation of the

    accumulation of damage as the solder material deforms

    plastically during the dynamic board flexing induced by drop

    impact (Fig. 2). The accumulated plastic strain criterion is

    attractive due to its convenience and simplicity, because it isable to convert complex loading sequences, such as the

    damping behaviour of a flexural strain waveform, shown in

    Fig. 3, into a single damage estimate.

    Yeh et. al. [9] explored the use of accumulated plastic

    strain in an FE drop impact study of the effects of kinematic

    and isotropic hardening on solder material response. The

    accumulated plastic strain results were used to explain the

    relative performance of joints at different locations within a

    single test vehicle. This relative comparison is similar to

    relative comparisons using peak interconnection stresses [6].

    PCB

    outermost solder

    joint

    package

    Shock table

    Test

    vehicle

    1-4244-0152-6/06/$20.00 2006 IEEE 1003 2006 Electronic Components and Technology Conference

  • 7/23/2019 4 point bend testing

    2/6

    Figure 3: Typical drop impact strain waveform

    indicating rapid damping of strain oscillations

    However, the FE studies using the concept of accumulated

    plastic strain have yet to establish a quantitative link, beyond

    relative comparisons, between plastic strain and failure. For

    example, it has yet to be proven if two different loading

    histories will result in failure when the same accumulated

    plastic strain is reached. Figure 4 illustrates this issue.

    Figure 4: Does failure occur at the same accumulated

    plastic strain for different loading histories?

    A very comprehensive test methodology and failure model

    for impact testing was introduced by Varghese and Dasgupta

    [13]. The proposed failure model is based on the calculation

    of a damage estimate which takes into account local PCB

    strain amplitudes, flexural modes and number of flexural

    cycles. Although the Paris Law for crack propagation is

    stated as the basis of the failure criterion, the assumptions

    used in the analysis result in a basic power law (e.g. Coffin-Manson) equation. The power law fatigue constants were

    deduced from impact experiments, based on several implicit

    assumptions which are discussed later in this paper. The

    complex dynamic responses of the PCB during impact were

    handled through 1) wavelet analysis to decompose the

    waveforms into their flexural mode components; and 2)

    fatigue cycle counting methods, similar to those

    recommended by fatigue test standards such as ASME

    E1049-85 [14], for determining the equivalent numbers of

    constant-amplitude cycles that can define the complex

    waveforms. A weightage factor is also applied to various

    modes to account for their different severities in causing joint

    damage. The damage estimate proposed is therefore the total

    amount of damage (be it crack initiation or growth)

    contributed by the various modes and amplitudes present in

    the drop impact waveforms. Lall et. al. [15] proposed a

    similar failure criterion involving wavelet analysis and fatigue

    cycle counting, but did not detail a treatment for various

    flexural modes.

    In view of the failure criteria described above, this paperintroduces high-speed bend testing methods for determining

    the constant-amplitude fatigue life of board assemblies rather

    than deducing them from drop impact tests. The results

    obtained will be able to provide insight into these failure

    criteria.

    1.3 Drop testing vs bend testing

    The use of bend tests instead of drop tests to study the

    drop impact reliability of PCB assemblies is motivated by

    earlier findings [16-17] showing that solder joint damage due

    to direct inertial loading is minimal. The studies found that

    the major cause of solder joint damage is the differential

    flexing between PCB and package, as shown in the FE plot of

    Fig. 2. In any case, the rapid motion of the PCB during adynamic bend test would already include some inertial

    loading. The JEDEC standard can in fact be viewed as a bend

    test method, although the bending is induced in an indirect

    manner using a shock pulse.

    The high frequencies and local accelerations of the PCB in

    drop impact pose challenges to performing displacement-

    controlled bend testing at drop impact strain rates. In a

    JEDEC board-level drop test, the impact pulse applied to the

    board assembly results in free vibration of the board at

    frequencies which ranging from 200 to 300 Hz. In product-

    level tests, the free vibration frequency of the PCB mounted

    within its housing has been found to be even higher [17-18],

    with frequencies of 400 to 500 Hz being common. Severalcases of higher flexural modes approach 1000 Hz. Given a

    frequency of 300 Hz and displacement amplitude of 3 mm for

    example, the accelerations at local points of the PCB can

    exceed 1000 Gs. The fastest hydraulic universal testers can

    achieve frequencies of at most 50 Hz at very low amplitudes.

    Electromagnetic shakers are unable to achieve large enough

    displacements and accelerations at the frequencies required.

    Hence the need for a new test equipment and method for

    performing high-speed bend testing.

    The JEDEC board-level drop test is also costly in terms of

    the time required to run the tests [7]. Several studies have

    even reported performing up to a hundred drops to induce

    failure. One of the reasons for having 15 packages on thestandard test vehicle is to generate several loading histories at

    different locations on the board in a single drop test.

    However, subsequent analysis is complicated as evident from

    the FE modeling studies done to model the JEDEC drop test.

    The ideal bend test is a four-point bend test on a single

    package board, provided it can be performed fairly quickly.

    Also, a displacement-controlled bend test and simple test

    dimensions may be consistent enough to eliminate the need

    for strain gauging.

    -4000

    -3000

    -2000

    -1000

    0

    1000

    2000

    3000

    4000

    0 100 200 300 400 500

    Time (ms)

    Strain

    (

    )

    0 0.02 0.04 0.06 0.08 0.1 0.12 0 0.02 0.04 0.06 0.08 0.1 0.12

    0 0.02 0.04 0.06 0.08 0.1 0.120 0.02 0.04 0.06 0.08 0.1 0.12

    Failure observed Predicted failure?

    1004 2006 Electronic Components and Technology Conference

  • 7/23/2019 4 point bend testing

    3/6

    Several novel bend test methods have been recently

    proposed for drop impact reliability testing. Reiff and Bradley

    [4] proposed a four-point bend test method in which a loading

    anvil is dropped onto a board assembly test vehicle, thus

    producing a bending pulse of a high frequency. The reported

    results showed a single half-sine bending pulse instead of the

    series of rapidly damping oscillations of drop impact. Their

    bend test method is similar in concept to the use of a

    pendulum impact test to induce bending in the PCB [13].However, the tests are velocity-controlled, and the flexing

    frequencies are largely dependent on both the anvil mass as

    the board stiffness. Chai et. al. [8] performed cyclic bend

    testing at frequencies of up to 20 Hz at the strain amplitudes

    of drop impact using a hydraulic universal tester. They

    concluded that because of the large differences in the

    frequencies (strain rates) of the cyclic tests and drop tests, no

    correlation was observed between the bend test and drop

    impact results.

    Figure 5: High-speed bend tester with 4-point

    bend test anvil

    Figure 6: Strain waveforms generated by high-speed

    bend tester using a sinusoidal cam

    The high-speed bend tester presented in this paper

    performs constant-amplitude displacement-controlled bendtesting at frequencies up to 500 Hz. Displacement control is

    possible through the use of a cam driven by a rotary motor.

    The cam enforces a particular displacement profile such as a

    sinusoidal or versine pulse. A lever connected to the cam

    follower controls the deflection amplitude. Fig. 5 shows the

    bend tester with four-point bend test anvil while Fig. 6 shows

    successive PCB strain waveforms generated by the tester.

    Each sinusoidal pulse is a single bending cycle. This

    equipment is estimated to achieve the equivalent bending

    cycles of 100 drops in several minutes. The tester is also able

    to perform single bending cycles at high speed, which is not

    possible in a drop test owing to the oscillations of the board.

    The relative performance of the various alloy/pad

    combinations in the high-speed bend tests has been found to

    be a match to their relative performance in JEDEC drop

    testing.

    2. Experimental procedure

    The test vehicles used were single-component board

    assemblies with solder mask defined (SMD) pads. The

    packages used were not actual IC packages, but PCB

    substrates with an identical pad finish to the boards, as shown

    in Fig. 7. The reason for using such dummy packages is to

    ensure that the joint qualities at the component and board

    sides are the same, and that the correct interface is being

    tested to failure. This is in view of several studies which have

    reported failures occurring at the package side [11,19]. Given

    similar board and component interfaces, fatigue failure should

    occur at the board side, according to earlier modeling studies

    performed [20].

    Figure 7: Board assembly test vehicle

    Figure 8: Pad layout showing corner joints and daisy

    chain circuitry

    The pad layout of the board is designed such that joint

    failures will occur at only the four corner interconnections

    shown in Fig. 8. This is because PCB bending results in the

    largest stresses occurring on the outermost joints (Fig. 2).

    These joints almost always fail first. FE modeling has shown

    that the failure of a corner joint has a minimal effect on rest of

    the corner joints. Also, due to symmetry, the stresses on thecorner joints are similar. The corner joints are essentially

    independent of one another and four independent data points

    can be obtained from one board sample. The board and

    components have circuitry for daisy chain monitoring, also

    seen in Fig. 8. In addition, each corner joint has its own

    resistance monitoring circuit. This design will aid failure

    analysis because the failure joint is known immediately, in

    contrast to a full array BGA package where the daisy chains

    may run through multiple solder joints. Failure is considered

    to be 10which is more than enough to represent a complete

    Loading anvil

    Board assembly

    -2000

    -1000

    0

    1000

    2000

    0 100 200 300

    Time (ms)

    PCBs

    train(

    )

    PCB substrate

    as component

    Corner pad

    1005 2006 Electronic Components and Technology Conference

  • 7/23/2019 4 point bend testing

    4/6

    500

    1000

    1500

    2000

    2500

    3000

    3500

    0 200 400 600 800

    Cycles to failure (N)

    PCBs

    train

    ()

    PbSn/ENIG 40 HzPbSn/ENIG 140HzPbSn/ENIG 250HzSnAg/ENIG 40HzSnAg/ENIG 250HzSnAgCu/ENIG 40Hz

    opening of a solder joint. The tracking of crack progression in

    each corner solder joint is also possible and is described in

    another study [21]. Other details on the board and component

    dimensions are shown below:

    Board/component dimensions

    Pitch: 1 mm

    Solder ball diameter: 0.3 mm

    SMD pad diameter: 0.25 mm

    Board thickness: 0.8 mm

    Component thickness: 0.8 mm

    I/O count: 40 (refer to Fig. 8 for exact design)

    To obtain data for constructing the power law fatigue

    curves, a series of high-speed four-point bend tests were

    performed for various frequencies, strain amplitudes and

    alloy/pad combinations. The following variables were used:

    Bend test variables

    Bending frequencies: 40 Hz, 140 Hz, 250 Hz

    PCB strains: Ranging from 1000 to3000

    Solder alloys: PbSn, SnAg, SnAgCu

    Pad finish: ENIG, Organic solder preservative (OSP)

    The four-point bend test setup had an inner span of 30 mm

    and an outer span of 90 mm, as shown in Fig. 9. A strain

    gauge was mounted at a specific location on the PCB adjacent

    to the package, as shown in Fig. 9. This will measure the

    reference strain which may be used in FE modeling. Using the

    high-speed bend tests, the number of cycles to failure for a

    specific frequency and amplitude is obtained for each corner

    solder joint. The failure curves of PCB strain vs. cycles-to-

    failure N (-N curves) were then plotted from the data

    collected.

    Figure 9: Four-point high-speed bend test setup

    3. Experimental results

    Figures 10 and 11 show the -N curves for ENIG and OSP

    pad finish respectively. Several trends may be seen. The lead-

    free solders fail in significantly fewer cycles than the PbSn

    solders on both ENIG and OSP pads. A decrease in reliability

    with frequency is seen for PbSn on both finishes and

    SnAg/OSP, but not for SnAg/ENIG. Tests were also done on

    two batches of PbSn/OSP boards to determine the effect of

    different processing parameters. Batch 1 can be seen to have

    a better reliability than Batch 2 of the PbSn/OSP boards.

    More tests will be done in the future to obtain statistically

    robust experimental curves. To obtain failure criteria based on

    solder materials, the PCB strain will also need to be

    converted to solder strain in FE modeling using accurate

    dynamic materials properties.

    Figure 10: -N curves for ENIG pad finish with various

    solder alloys and test frequencies

    Figure 11: -N curves for OSP pad finish with various

    solder alloys and test frequencies

    Figure 12: Crack through bulk solder material transitioning to

    intermetallic failure (SnAg/ENIG, 250Hz, 1500strain)

    PCB

    ComponentStrain

    gauge Sliding support

    30 mm

    90 mm

    1500

    1700

    1900

    2100

    2300

    2500

    2700

    2900

    3100

    3300

    3500

    0 200 400 600 800

    Cycles to failure (N)

    PCBs

    train

    ()

    PbSn/OSP - Batch 1 - 40HzPbSn/OSP - Batch 1 - 140HzPbSn/OSP - Batch 2 - 40HzPbSn/OSP - Batch 2 - 140HzSnAg/OSP 40HzSnAg/OSP 140HzSAC/OSP 40Hz

    1006 2006 Electronic Components and Technology Conference

  • 7/23/2019 4 point bend testing

    5/6

    The high-bend tests produce similar failure modes as drop

    impact, with various mixes of bulk and intermetallic failure.

    Figures 12 and 13 show the cross-sections of two joints from

    the same board with different failure modes.

    Figure 13: Almost completely intermetallic failure

    (SnAg/ENIG, 250Hz, 1500)

    4. Discussion on failure criteria

    The damage estimate methodologies [13,15] which

    deduce the power law fatigue constants from drop testexperiments depend on two important assumptions:

    1) A linear damage rule (also known as Miners rule) [22] is

    assumed when calculating the damage contributed by the

    various amplitudes. For example, given that the fatigue life

    at a certain strain amplitude S is N cycles, each cycle at

    amplitude S will consume 1/N of the fatigue life of the

    joint. In a load history consisting of various amplitudes, it

    is then assumed that the same linear damage rule is applied

    for all amplitudes.

    2) The effect of load sequence is ignored. For example, it is

    assumed that the damage estimate is the same whether

    testing is conducted at A) strain S1for N1cycles followed

    by strain S2 for N2 cycles or; B) strain S2 for N2 cyclesfollowed by strain S1 for N1cycles. This assumption may

    not be a critical issue given that most board-level tests

    consist of a series of similar drops. It may be an issue if the

    failure criteria is to have a more general application

    beyond specific board level test setups.

    These studies also reported that validation of the

    calculated damage estimates was performed by conducting

    impact tests at different impact orientations or by validating

    different board locations other than those used to derive the

    power law constants. Based on the reported information, it is

    uncertain if the strain waveforms of the validation tests were

    sufficiently dissimilar to be able to independently verify

    whether the above assumptions hold true. If the validationwaveforms are too similar to the waveforms used to derive the

    constants, the damage criteria might be applicable only to that

    specific type of waveform and the specific test setup. The

    failure data obtained from high-speed bend tests will be able

    to verify if the above assumptions are reasonable.

    If accumulated plastic strain obtained from modeling is to

    be used as an accurate damage estimate, the material

    characteristics of the solder would have to be such that the

    failure criterion from modeling predicts the failure data of the

    experimental -N curves. For example, a purely kinematic

    hardening plastic model will not work because the implied

    linear relation between accumulated damage and number of

    plastic strain cycles results in a power law fatigue exponent of

    1. Even after conversion of solder plastic strain to PCB strain,

    the exponent for a kinematic model will not reach the

    exponent values of the experimental -N curves which range

    from 0.12 to 0.75 (Figs. 10 and 11). Also, if the plastic strain

    criterion is to have general applications beyond board-level

    drop, the material models used would have to incorporate aproper damage rule (linear or non-linear) and the effect of

    amplitude sequence (if any).

    It should noted that the -N curves of Figs. 10 and 11 do

    not distinguish between the stages of failure, be it crack

    initiation or crack propagation. For example, a high strain

    amplitude may result in immediate crack propagation while a

    low strain amplitude may have a relatively much longer

    initiation phase, but both are used to construct the same -N

    curve. The -N curves also do not distinguish between failure

    modes, such as bulk failure or intermetallic failure. There is

    still uncertainty as to whether basic power law fatigue curves

    can be applied across all failure phenomena encountered in

    drop impact. The fatigue failure in drop impact is much closerto very low cycle fatigue (VLCF) than low cycle fatigue

    (LCF). A model for solder based purely on strain range might

    not work in the VLCF regime.

    The failure data in this paper is presented as PCB strain

    vs number of cycles to failure N. A proper failure criteria

    would ideally be materials-based. This will involve

    converting PCB strain to solder strain using accurate dynamic

    solder material properties and taking into account the details

    of the solder joint structure.

    The -N failure curves only represent constant-amplitude

    failure data. Future work in this study will focus on 1) how

    damage is accumulated in a complex drop impact response; 2)

    the effect of sequence; and 3) the combined effect of variousfrequencies.

    5. Future challenges

    Much work has yet to be done to develop a practical and

    sufficiently accurate failure criterion which can have general

    applications other than board-level impact tests. Among the

    issues to be resolved are:

    1) Determining the damage rule to use in fatigue damage

    estimates.

    2) Assessing the effect of amplitude sequence in fatigue

    damage.

    3) Obtaining a full range of dynamic solder material

    properties since a solder joint in drop impact can

    experience a wide range of strain rates.

    4) Determining the effect of failure mode and crack

    progression on the damage estimate.

    5) Confirming or determining the fatigue model to use in the

    VLCF regime.

    6. Conclusions

    A high-speed bend test method has been introduced for

    more controllable and effective testing of drop impact

    reliability. The test method is being used to obtain failure

    criteria for drop impact.

    1007 2006 Electronic Components and Technology Conference

  • 7/23/2019 4 point bend testing

    6/6

    The development of a practical failure criteria for drop

    impact is a very challenging but valuable area of study.

    Various groups have proposed several interesting drop impact

    failure criteria and test methodologies. The current methods

    for failure prediction in drop impact are:

    1) Number of drops to failure in the JEDEC drop test, where

    comparisons of drops-to-failure are heavily dependent on

    the test setup.

    2) An accumulated plastic strain failure criterion.3) Damage estimate failure criteria based on drop impact

    waveforms and power law fatigue equations.

    7. References

    1. Anand Lal, Edwin Bradley, and Jignesh Sharda, Effect of

    Reflow Profiles on the Board Level Drop Reliability of

    Pb-Free (SnAgCu) BGA Assemblies, Proc 55th

    Electronic Components and Technology Conference,

    2005, pp. 945-953.

    2. Keith Newman, BGA Brittle Fracture Alternative

    Solder Joint Integrity Test Methods, Proc 55th

    Electronic Components and Technology Conference,

    2005, pp. 1194-1201.

    3. Chang-Lin Yeh, Yi-Shao Lai, Hsiao-Chuan Chang, and

    Tsan-Hsien Chen, Correlation Between Package-level

    Ball Impact Test and Board-level Drop Test, Proc 7th

    Electronics Packaging Technology Conference, 2005, pp.

    270-275.

    4. Dave Reiff, and Edwin Bradley, A Novel Mechanical

    Shock Test Method to Evaluate Lead-Free BGA Solder

    Joint Reliability, Proc 55th Electronic Components and

    Technology Conference, 2005, pp. 1519-1525.

    5. JEDEC Solid State Technology Association, JESD22-

    B111: Board Level Drop Test Method of Components for

    Handheld Electronic Products , July 2003

    6. Ahmer Syed, Seung Mo Kim, Wei Lin, Jin Young Khim,

    Eun Sook Song, Jae Hyeon Shin, and Tony Panczak, A

    Methodology for Drop Performance Prediction and

    Application for Design Optimization of Chip Scale

    Packages, Proc 55th Electronic Components and

    Technology Conference, 2005, pp. 472-479.

    7. Steve Groothuis, Changming Chen, and Radovan

    Kovacevic, Parametric Investigation of Dynamic

    Behavior of FBGA Solder Joints in Board-Level Drop

    Simulation, Proc 55th Electronic Components and

    Technology Conference, 2005, pp. 499-503.

    8. Chai T.C., Sharon Quek, Hnin W.Y., Wong E.H.,Julian

    Chia, Wang Y.Y., Tan Y.M., and Lim C.T., Board Level

    Drop Test Reliability of IC Packages, Proc 55th

    Electronic Components and Technology Conference,2005, pp. 630-636.

    9. Chang-Lin Yeh, Yi-Shao Lai, and Chin-Li Kao,

    Prediction of Board-level Reliability of Chip-scale

    Packages Under Consecutive Drops, Proc 7th

    Electronics Packaging Technology Conference, 2005, pp.

    73-80.

    10. Jing-en Luan, and Tong Yan Tee, Effect of Impact Pulse

    Parameters on Consistency of Board Level Drop Test and

    Dynamic Responses, Proc 55th Electronic Components

    and Technology Conference, 2005, pp. 665-673.

    11. Pekka Marjamki, Toni Mattila, and Jorma Kivilahti,

    Finite Element Analysis of Lead-Free Drop Test

    Boards, Proc 55th Electronic Components and

    Technology Conference, 2005, pp. 462-466.

    12. Liping Zhu, and Rick Williams, Plastic Strain Based

    Criterion for Reliability Assessment of CSP Packages

    Subjected to Dynamic Shock Loads, Proc ASMEInterPACK 05, 2005, IPACK2005-73059, pp. 1-8.

    13. J. Varghese, and A. Dasgupta, Test Methodology for

    Impact Testing of Portable Electronic Products, Proc

    2003 ASME International Mechanical Engineering

    Congress & Exposition, IMECE2003-41844, pp. 1-7.

    14. American Society for Testing and Materials (ASTM),

    E1049-85(2005): Standard Practices for Cycle Counting

    in Fatigue Analysis, 2005.

    15. Pradeep Lall, Dhananjay Panchagade, Prakriti Choudhary,

    Jeff Suhling, and Sameep Gupte, Failure-Envelope

    Approach to Modeling Shock and Vibration Survivability

    of Electronic and MEMS Packaging, Proc 55th

    Electronic Components and Technology Conference,2005, pp. 480-490.

    16. E.H. Wong, K.M. Lim, N. Lee, S. Seah, C. Koh, and J.

    Wang, Drop Impact Test Mechanics and Physics of

    Failure, Proc. 4th Electronic Packaging Technology

    Conference, 2002, pp. 327-333.

    17. Seah S.K.W., Wong E.H., Ranjan R.,Lim C.T., Mai Y.W.,

    Understanding and Testing for Drop Impact Failure,

    Proc ASME InterPACK 05, 2005, IPACK2005-73047, pp

    1-6.

    18. Lim C.T., Wong E.H., Ang C.W., Tan L.B., Tan V.B.C.,

    and Shim V.P.W., "Drop Impact Study of Handheld

    Electronic Products," Proc 5th International Symposium

    on Impact Engineering, UK, 2004.19. Mikko Alajoki, Luu Nguyen, and Jorma Kivilahti, Drop

    Test Reliability of Wafer Level Chip Scale Packages,

    Proc 55th Electronic Components and Technology

    Conference, 2005, pp. 637-644.

    20. E. H. Wong, Y-W Mai, and S. K. W. Seah, Board Level

    Drop Impact Fundamental and Parametric Analysis,

    ASME Journal of Electronic Packaging, 2005, Volume

    127, Issue 4, pp. 496-502.

    21. Seah S.K.W., Wong E.H., Mai Y.W., Rajoo R., and Lim

    C.T. Failure Mechanisms of Interconnections in Drop

    Impact, Proc 56th Electronic Components and

    Technology Conference, 2006.

    21. Bannantine J.A., Comer J.J., and Handrock J.L.,Fundamentals of Metal Fatigue Analysis, 1989, Prentice-

    Hall.

    1008 2006 Electronic Components and Technology Conference