Microelectronics Reliability



Similar documents
CHAPTER 3 MODAL ANALYSIS OF A PRINTED CIRCUIT BOARD

COMPUTATIONAL ENGINEERING OF FINITE ELEMENT MODELLING FOR AUTOMOTIVE APPLICATION USING ABAQUS

Soil Dynamics Prof. Deepankar Choudhury Department of Civil Engineering Indian Institute of Technology, Bombay

RANDOM VIBRATION AN OVERVIEW by Barry Controls, Hopkinton, MA

THE COMPOSITE DISC - A NEW JOINT FOR HIGH POWER DRIVESHAFTS

Comparison of the Response of a Simple Structure to Single Axis and Multiple Axis Random Vibration Inputs

Precision Miniature Load Cell. Models 8431, 8432 with Overload Protection

Back to Elements - Tetrahedra vs. Hexahedra

Abaqus Technology Brief. Automobile Roof Crush Analysis with Abaqus

Structural Integrity Analysis

Methods to predict fatigue in CubeSat structures and mechanisms

Tower Cross Arm Numerical Analysis

4 SENSORS. Example. A force of 1 N is exerted on a PZT5A disc of diameter 10 mm and thickness 1 mm. The resulting mechanical stress is:

Solved with COMSOL Multiphysics 4.3

DYNAMIC ANALYSIS ON STEEL FIBRE

Optimum proportions for the design of suspension bridge

Herschel PACS Test Procedure STM PACS-KT-PR-011 Issue: 1 Revision: -

Fluid structure interaction of a vibrating circular plate in a bounded fluid volume: simulation and experiment

METHODS FOR ACHIEVEMENT UNIFORM STRESSES DISTRIBUTION UNDER THE FOUNDATION

ANALYTICAL AND EXPERIMENTAL EVALUATION OF SPRING BACK EFFECTS IN A TYPICAL COLD ROLLED SHEET

CRASH ANALYSIS OF AN IMPACT ATTENUATOR FOR RACING CAR IN SANDWICH MATERIAL

The Basics of FEA Procedure

DYNAMICAL ANALYSIS OF SILO SURFACE CLEANING ROBOT USING FINITE ELEMENT METHOD

International Journal of Engineering Research-Online A Peer Reviewed International Journal Articles available online

The elements used in commercial codes can be classified in two basic categories:

AIR RESONANCE IN A PLASTIC BOTTLE Darrell Megli, Emeritus Professor of Physics, University of Evansville, Evansville, IN dm37@evansville.

In-situ Load Testing to Evaluate New Repair Techniques

White Paper. Recommendations for Installing Flash LEDs on Flex Circuits. By Shereen Lim. Abstract. What is a Flex Circuit?

Introduction to Solid Modeling Using SolidWorks 2012 SolidWorks Simulation Tutorial Page 1

8.2 Elastic Strain Energy

Clutch and Operation as a System

Bending, Forming and Flexing Printed Circuits

MASTER DEGREE PROJECT

Similar benefits are also derived through modal testing of other space structures.

The Calculation of G rms

Numerical simulation of ground impact after airdrop

EXPERIMENTAL ANALYSIS OF CIRCULAR DISC WITH DIAMETRAL SLOTS

Dispersion diagrams of a water-loaded cylindrical shell obtained from the structural and acoustic responses of the sensor array along the shell

Design Validation and Improvement Study of HVAC Plumbing Line Assembly under Random Loading Condition

BLIND TEST ON DAMAGE DETECTION OF A STEEL FRAME STRUCTURE

DEVELOPMENT AND APPLICATIONS OF TUNED/HYBRID MASS DAMPERS USING MULTI-STAGE RUBBER BEARINGS FOR VIBRATION CONTROL OF STRUCTURES

SIESMIC SLOSHING IN CYLINDRICAL TANKS WITH FLEXIBLE BAFFLES

DYNAMIC ANALYSIS OF THICK PLATES SUBJECTED TO EARTQUAKE

Nonlinear analysis and form-finding in GSA Training Course

PERPLEXING VARIABLE FREQUENCY DRIVE VIBRATION PROBLEMS. Brian Howes 1

Force measurement. Forces VECTORIAL ISSUES ACTION ET RÉACTION ISOSTATISM

Sensor Performance Metrics

Tensile fracture analysis of blunt notched PMMA specimens by means of the Strain Energy Density

The accelerometer designed and realized so far is intended for an. aerospace application. Detailed testing and analysis needs to be

Estimating Acoustic Performance of a Cell Phone Speaker Using Abaqus

A Case Study Comparing Two Approaches for Applying Area Loads: Tributary Area Loads vs Shell Pressure Loads

Optiffuser. High-performance, high bandwidth lightweight 1D diffuser.

GMC 2013: Piping Misalignment and Vibration Related Fatigue Failures

An Overview of the Finite Element Analysis

Vibrations of a Free-Free Beam

New approaches in Eurocode 3 efficient global structural design

The Fundamental Principles of Composite Material Stiffness Predictions. David Richardson

DESIGN OF SLABS. 3) Based on support or boundary condition: Simply supported, Cantilever slab,

INJECTION MOLDING COOLING TIME REDUCTION AND THERMAL STRESS ANALYSIS

Rotation: Moment of Inertia and Torque

Physics 9e/Cutnell. correlated to the. College Board AP Physics 1 Course Objectives

Finite Element Formulation for Plates - Handout 3 -

Subminiature Load Cell Model 8417

Objectives. Experimentally determine the yield strength, tensile strength, and modules of elasticity and ductility of given materials.

Safakcan Tuncdemir 1, William M. Bradley *2. 1. Introduction

Structural Axial, Shear and Bending Moments

B.TECH. (AEROSPACE ENGINEERING) PROGRAMME (BTAE) Term-End Examination December, 2011 BAS-010 : MACHINE DESIGN

PROPERTIES OF MATERIALS

Cylinder Head Gasket Contact Pressure Simulation for a Hermetic Compressor

Tutorial for Assignment #2 Gantry Crane Analysis By ANSYS (Mechanical APDL) V.13.0

How To Write An Analysis System For Bridge Test

Mounting instructions. Acceleration Transducer B12. B 26.B12.10 en

Finite Element Formulation for Beams - Handout 2 -

INTER-NOISE AUGUST 2007 ISTANBUL, TURKEY

WAVEGUIDE-COAXIAL LINE TRANSITIONS

METHOD OF STATEMENT FOR STATIC LOADING TEST

Simulation in design of high performance machine tools

MSC/SuperModel A CAE Data Management and Advanced Structural Modeling System

Rock Bolt Condition Monitoring Using Ultrasonic Guided Waves

Material Optimization and Weight Reduction of Drive Shaft Using Composite Material

CASE STUDY. University of Windsor, Canada Modal Analysis and PULSE Reflex. Automotive Audio Systems PULSE Software including PULSE Reflex, Transducers

DESIGN AND ANALYSIS OF BRIDGE WITH TWO ENDS FIXED ON VERTICAL WALL USING FINITE ELEMENT ANALYSIS

Lap Fillet Weld Calculations and FEA Techniques

Calibration and Use of a Strain-Gage-Instrumented Beam: Density Determination and Weight-Flow-Rate Measurement

Determination of g using a spring

Numerical Analysis of Independent Wire Strand Core (IWSC) Wire Rope

Spring Force Constant Determination as a Learning Tool for Graphing and Modeling

Flex Circuit Design and Manufacture.

Finite Element Model Calibration of a Full Scale Bridge Using Measured Frequency Response Functions

An Introduction to Rigid-Flex PCB Design Best Practices

THE DETERMINATION OF DELAMINATION STRAIN ENERGY RELEASE RATE OF COMPOSITE BI-MATERIAL INTERFACE

AP1 Oscillations. 1. Which of the following statements about a spring-block oscillator in simple harmonic motion about its equilibrium point is false?

STUDY OF DAM-RESERVOIR DYNAMIC INTERACTION USING VIBRATION TESTS ON A PHYSICAL MODEL

FXA UNIT G484 Module Simple Harmonic Oscillations 11. frequency of the applied = natural frequency of the

DfRSoft 2-Day Course Shock & Vibration: Test, Design, and Design Assurance

DAMPING BEHAVIOUR OF CAST AND SINTERED ALUMINIUM

Advanced bolt assessment through process automation

Transducer Mounting and Test Setup Configurations. Rick Bono The Modal Shop

Pancake-type collapse energy absorption mechanisms and their influence on the final outcome (complete version)

Free vibration of CLT plates

Transcription:

Microelectronics Reliability 50 (2010) 86 97 Contents lists available at ScienceDirect Microelectronics Reliability journal homepage: www.elsevier.com/locate/microrel Accuracy of simplified printed circuit board finite element models Robin Alastair Amy a, *, Guglielmo S. Aglietti a, Guy Richardson b a Astronautical Research Group, School of Engineering Sciences, University of Southampton, UK b Surrey Satellite Technology Limited, Guildford, Surrey, UK article info abstract Article history: Received 14 October 2008 Received in revised form 23 February 2009 Available online 17 October 2009 Electronic components are prone to failure due to shock or vibration loads. To predict when this failure may occur it is necessary to calculate the vibration response of the printed circuit board (PCB); this is most usually achieved through use of simplified finite element (FE) models. The accuracy of these FE models will be mainly dependant on various sources of error, including: manufacturing variability, which will cause supposedly identical printed circuit boards to behave differently (including variability in materials and assembly, as well as dimensional tolerances); inaccuracy in the model input parameters, which is caused by either the modelling assumptions used or poor measurement technique; and errors in the solution process (e.g. linear solutions in non-linear situations). This paper investigates experimentally the contribution of these effects, this is achieved by first looking at measurement of input parameters and to what accuracy a PCB can reasonably be modelled, and then secondly measuring the extent of manufacturing and assembly induced variability. When these contributions have been defined, it will be possible to assess the confidence in any FE PCB model. Ó 2009 Elsevier Ltd. All rights reserved. 1. Introduction Shock and vibration loads imposed on a printed circuit board may, if severe enough, cause the attached electronic components to fail. The probability of this failure occurring is strongly influenced by the local PCB vibration response; therefore, if equipment vibration reliability is to be predicted then so too must the PCB local vibration response. The most convenient and commonly used method of predicting the local vibration response is through the creation of finite element models. Many good examples of detailed finite element models already exist in the current literature [1 8], where the complexity of these models is justified as the component internal stresses are required; however, the long time required to build and solve the models cannot be justified when only the PCB response is required. It is possible to greatly simplify the creation of these models by using lead spring constants in place of detailed lead models [9 11]. Further simplification can be achieved by removing the detailed 3D meshes of the component from the FE model altogether: instead their effects can be included by artificially increasing the stiffness and density of the underlying PCB properties in the FE model [12,13]. These artificial properties can either be included at the exact component location or averaged (sometimes called smearing ) over the entire area of the board. As well as the detailed and smeared models there are also models that use simple block 3D elements to recreate component effects [4,14 16], these models * Corresponding author. E-mail address: robinamy@soton.ac.uk (R.A. Amy). are especially useful when relatively tall or big components are present. In addition to all the work on accurately incorporating component effects into the model, other work has looked the importance of accurately incorporating boundary conditions into the model [17,18], or specifically considering the effect of boundary condition modification with respect to reducing the vibration response [18 21]. Finally, some previous research has noted the importance of considering non-linear effects during modelling [22,23]. It is shown in this work that the first major difficulty in creating good PCB FE models is in specifying the input parameters, namely: stiffness, density, damping and boundary conditions. The accuracy with which each of these parameters is specified will determine the accuracy of the response predicted by the final model. Any inaccuracy in the input parameters will occur due to either poor initial measurements or, when measurements are not possible, poor estimation of these factors. A secondary difficulty in creating accurate PCB FE models is due to variation that may occur in each of the input parameters, where the variation will cause supposedly identically manufactured and assembled PCBs to have different responses. Finally, the model accuracy will be limited by parameters and sources of variation that cannot be easily specified or modelled. These may include non-linear effects, limitations of the FE mesh and other effects too complicated to be easily included in the model (i.e. air or acoustic influences). Most often these are accounted for by including appropriate safety factors. Using an experimental approach and keeping in mind the three points above, this paper will assess how accurately simplified mod- 0026-2714/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.microrel.2009.09.001

R.A. Amy et al. / Microelectronics Reliability 50 (2010) 86 97 87 els can predict a given PCB s response. The approach will involve two stages: the first stage will look at the creation of idealised theoretical FE models of PCBs and how accurately these models predict the actual response. The second stage will quantify the variations that occur due to small differences in manufacturing and assembly, allowing quantification of the expected variability and its effect on response. There are three main contributions of this work: (1) future researchers in this area can follow a similar process to the one shown here and obtain expected accuracy values specific to their own equipment and manufacturing techniques, (2) the observations presented here allows the modelling effort spent on future PCB models to be used much more effectively, and (3) the expected accuracies published here can be used as ball-park starting values for future work. 2. Current practice The whole FEA process can be broken into three main parts: (1) modelling (including estimating input parameters, creating a mesh and incorporating measured parameters in the model), (2) analysis (type of solution, linear or non-linear, etc), and (3) post-processing (looking at appropriate response parameters, such as: acceleration, curvature or deflection). Each of these three parts will be considered in turn. 2.1. Creating FE models of PCBs In creating any FE model of a PCB it is convenient to break the model down into five separate areas: PCB properties, Component effects, chassis, damping and boundary conditions. 2.1.1. PCB properties The main difficulty in creating a model of a PCB is not in creating the mesh but in specifying the properties: stiffness moduli, poisson ratio, density and thickness. These properties may be provided by the PCB manufacturer but those that are not should be measured. The Young s modulus is most simply calculated using a static bend test [13]. It is important to note that as most PCBs are laminates their properties may vary depending on the direction of loading, this necessitates that the Young s modulus be measured in both the x and y axis of the PCB (where axis are defined as in Fig. 1). The shear modulus of a PCB may most conveniently be determined through use of a static four point bend test [13] as shown in Fig. 1, allowing the shear modulus to be calculated using: G xy ¼ 3K tab 4t 3 where K t is the slope of the load displacement curve, a and b are the specimen edge lengths and t is the specimen thickness. Fig. 1. Diagram of experimental set-up of torsion test. Loads are placed on the four corners of the specimen and the corner deflection is measured. Total load = 2F. ð1þ The density can easily be found if the PCB mass, width, length and thickness are known, where there is the implicit assumption that the thickness is constant over the PCB. Additionally, removal of the PCB s copper surface during manufacturing will reduce both the mass and stiffness. Because of this fact it is important that any bend testing should be performed on the etched PCB, as this reduces possible errors. 2.1.2. Component effects The following brief discussion on component modelling is only included to provide the reader with the minimum necessary comprehension of the subject. This subject is too broad to satisfactorily cover in this work; therefore, whilst reading this work, it should be remembered that component modelling is not within its scope. When components are soldered to a PCB they will locally increase the mass and stiffness of the PCB. For good accuracy, especially when large numbers of components are present, the FE model should include this added mass and stiffness effect. Theoretically, this could be achieved by including each individual component in the model. Several good examples of such detailed models exist [1 4], the high level of detail in these models is justified as these works look at stresses within the component. However, if only the PCB response is required then detailed models require too much time and effort then can normally be justified. One solution that avoids excessively complicated models is to assume that the additional mass and stiffness of the component can be included by artificially increasing the PCB Young s modulus and density [12,13]. This increase may be simplified by averaging (or smearing ) the additional mass and stiffness over the entire area of the board. The accuracy of the method depends on the level of simplification used and the mass and stiffness of the components present [24]. As a general rule, small components (discrete resistors and capacitors) may usually be ignored whilst larger and heavier components (heat-sinks, power transformers, large capacitors, etc) should always be modelled. The method of including the larger components depends on the information required, i.e. if only the PCB response is required then usually either smeared or simple 3D models are sufficient, whilst if component lead stresses are required then detailed 3D models are necessary. However, these are only very general rules, and ideally the strengths and weaknesses of each method should be understood and considered each time they are utilised. 2.1.3. Chassis A general rule of thumb for the FE modelling of any piece of equipment is to always model the next level up from the object of interest [25]. In the case of electronic equipment that means the chassis (sometimes referred to as enclosure) should also be modelled. Failure to include the chassis response in the model may severely affect the accuracy, unless the chassis is extremely rigid relative to the PCB. Generally, modelling a chassis is fairly simple as they are relatively straightforward structures, although it is always highly recommended to validate the FE model predicted response of the bare chassis with experimentally measured values. 2.1.4. PCB boundary conditions Once the FE models of the PCB and chassis have been created, the next step is to combine the two together so that the overall response can be calculated. To achieve this, rigid and/or spring FE elements are used to connect the two models, where these elements are intended to represent the effect of the PCB-chassis fixing method (e.g. bolts or card-lok systems). In terms of translational displacement, most fixings will be stiff enough that they will effectively rigidly constrain the PCB translational displacement to that of the chassis (this is proved in Section 4 of this work); this means that simple rigid FE elements are suffi-

88 R.A. Amy et al. / Microelectronics Reliability 50 (2010) 86 97 cient to constrain the x, y and z displacement of the PCB to the chassis. However, in terms of rotational displacement all fixing methods will display some flexibility, therefore rotational spring elements are required to tie the two models together. The main difficulty in incorporating these rotational spring elements into the FE model is in specifying the value of the spring stiffness. There are two possible approaches to this based on whether a prototype of the PCB exists or not: first, if a prototype of the PCB and chassis does not exist then the users options are very limited and the only way to obtain spring constants is to either estimate the value based on subjective experience of previous fixings or to create a detailed FE model of the joint (which would probably be too time consuming to be practical). Secondly, if a real example of the PCB and chassis does exist then the accuracy may be improved, as experiments may be performed on this combined structure to calculate the rotational stiffness of that fixing [17]. This second approach requires an experimental set-up incorporating the fixing method of interest attached to a PCB to be created. It is then possible to use a trial and error ( tuning ) approach on an FE model of this structure, where the spring rotational stiffness tuned until the predicted frequencies match those measured experimentally. When the two frequencies match, assuming everything else in the model is correct, the stiffness used in the model is assumed the same as the real stiffness. This method has been illustrated in works by various authors [17,18], and also extended to allow calculation of a non-dimensional parameter for the rotational stiffness [17]. It is also pertinent to mention that previous studies have attempted to calculate the boundary rotational stiffness based on the static deflection of an experimental set-up, but the process was deemed impracticable [17]. The method described in this previous work [17] was originally intended for use with card-lok style fixing mechanisms, which provide clamping force along the entire edge of a PCB, not just in discrete location as happens with bolted PCBs. Thus, the method should be used with caution if it is applied to bolted-down PCBs, as it may not be a correct application of the method. If a large number of bolts are present on the edge of the PCB then the situation may be considered similar to that of a card-lok fastened PCB, whereas if the PCB is fixed in only a few locations such an assumption could prove unrealistic. Other works relevant to boundary conditions also exist. The effects of very small variations in screw tightness on shock response have been examined [26], showing that even half an M3 screw pitch of variation can alter the PCB response. The same work also showed that the number and location of fixings dramatically alters the response, similar results have also been shown in other work [27]. 2.1.5. Damping Although there are several techniques to experimentally measure the damping, in this paper we shall focus on the three most convenient methods: the logarithmic decrement method, the magnification-factor method and the bandwidth method [28]. The logarithmic decrement method is a time domain approach that calculates damping based on the free decay of oscillation in a structure. The following Eqs. (2) and (3) are used to calculate the damping (f ld logarithmic decrement). r ¼ 1 r ln x n ¼ 2pf ld qffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð2þ x nþr 1 f 2 ld 1 f ld ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð3þ 1 þð2p=rþ 2 where r is the logarithmic decrement per unit cycle, x is the amplitude of the structures response and r is the number of cycles apart in the time history. The bandwidth method is a frequency domain method that calculates damping from the frequency response of the PCB. Once the frequency response has been obtained, the response at the half power points is measured and the damping (f b bandwidth) calculated using Eq. (4). f b ¼ 1 2 Dx x r where Dx is the width of the frequency response curve at the half power level, and x r is the resonant frequency. The Magnification-factor method is also a frequency domain method where the damping (f mf magnification factor) is calculated from the peak transmissibility at resonance. Q ¼ 2f mf 1 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 f 2 mf or for low damping ðf < 0:1Þ this simplifies to Q ¼ 1 2f mf which is simply re-arranged to f mf ¼ 1 2Q When choosing which of the three methods to use the following points should be considered [28]. The frequency domain methods are poor for low damping (<1%) as the curves will be difficult to measure accurately due to the high rate of change of the frequency response curve. When such low damping is present then the time domain methods are preferred. Another point to consider is that the damping may increase with increasing deflection in the structure, necessitating the damping to be measured at more than one level of input vibration. Finally, all the methods here assume that only one mode is being excited, if other modes exist close to the one of interest then more detailed analysis is required outside the scope of this paper (see [28] for more information). Analytical methods of finding the damping value of a structure also exist. Steinberg [29] states that the transmissibility at resonance of an electronic sub-assembly is equal to two times the square root of the resonant frequency: p Q peak ¼ a ffiffiffiffiffiffi x r ð8þ where a is a fitting factor based on x r and Q peak is the transmissibility at resonance. The factor a equals 0.5 if ðx r 6 100Þ, 0.75 if ð100 < x r 6 200Þ, 1ifð200 < x r 6 400Þ and 2 if ð400 < x r Þ. Subsequently the damping can be calculated from the transmissibility by Eq. (7), provided that the level of damping is low ðf < 0:1Þ. Unfortunately the data on which Steinbergs method is based is unavailable and therefore the method is unverifiable. 2.2. Analysis stage Typically, the dynamic response of the board will be calculated by first performing a modal analysis and then a frequency response analysis using a mode superposition method. To avoid large errors when solving for the dynamic response of a PCB, two points should be considered [23]: first, the modal solution of the model should consider enough modes so that a significant fraction (roughly at least 90%) of the total mass of the structure is excited. In addition, when the board deflections are comparable to that of the board thickness a non-linear analysis is preferred [23]; however, no such analysis can be found in the current literature, most likely due to the difficulty in specifying the non-linear input parameters (e.g. non-linear edge rotational stiffness, frequency). ð4þ ð5þ ð6þ ð7þ

R.A. Amy et al. / Microelectronics Reliability 50 (2010) 86 97 89 In this work, model optimisation in terms of computer resources is not considered as the solutions times are very short, even on relatively modest hardware (<30 s). It is assumed, therefore, that any optimisation efforts would consume more time than is saved. However, more computer intensive modelling strategies (detailed 3D modelling of components for example) might benefit from such optimisation, in this case sub-structuring methods are most commonly used [12,13]. 2.3. Post-processing stage The final stage of the FE process is to convert the FEA output to obtain suitable output parameters; this depends on the purpose of the analysis. Most users will wish to compare the results with some pre-defined failure criteria; for example, relating the acceleration experienced by a component to probable time for it to fail. These failure criteria may take the form of local acceleration experienced by a component [30], local bending moments [31], local board surface strain [5,6] or board deflection [29]. The choice of which failure criteria to use depends on many factors, including: component size, component mass, mounting technology, manufacturing quality and board thickness. The abundance of these influencing factors makes it very difficult to recommend using one failure criteria over another. As general rule, however, the failure of lightweight components (e.g. QFP or BGA) correlates most strongly with either board curvature, bending moments or board deflection (the choice between which of these criteria to use is mainly based on ease of measurement). This is in contrast to heavy components, these are increasingly likely to fail from high accelerations as their mass increases, and this can be attributed to the difference in inertia between light and heavy components. It is important to remember that the choice of failure criteria requires a thorough understanding of the previous research and should not be taken lightly, for the strengths and weaknesses of each criteria should be fully considered, whilst a criteria that is correct for one situation may be incorrect for another. The difficulty in recommending specific failure criteria in no way discredits the work here, this is because the models created in this work are able to create any of these metrics with ease. 2.4. Additional software methods It is relevant to note that a commercial software package CalcePWA also exists for the calculation of the PCB response [2,32]. This software automates the creation of PCB FE models through a GUI interface. The modelling approach used in such software is not known as the software is proprietary, although accurate boundary condition modelling is assumed to be incorporated. and it is reported here to illustrate and discuss some typical PCBs response. 3. Properties determination In the first part of this section, the material and damping properties of two typical PCB set-ups will be measured, and then the results from these tests are to be used to create FE models of the PCBs. An analysis of manufacturing and assembly variability will be made after this comparison. Throughout all the tests in this work, the frequency response was measured with a dynamic signal Acquisition Board (NI PCI- 4472) and several small accelerometers attached to the boards (Piezotronics 0.6 g or 0.2 g). Where multiple tests were performed on identical boards the accelerometers exact position was ensured to be identical between each test. The two PCBs to be modelled will be hereafter referred to as Setup A and Set-up B, they consist of the following: Set-up A consists of an unpopulated PCB attached to an aluminium enclosure as shown in Fig. 2. The PCB is made of FR4 laminate and is attached to the enclosure with M2.5 bolts, it measures 289 mm by 316 mm and weighs 359 g. The enclosure has cross-members as shown in Fig. 10 (also known as an antivibration frame) to provide extra support for the centre of the PCB, it was attached to the shaker head expander with 8 M5 bolts and using 30 mm stand-offs to minimise air-pumping effects. In later tests the PCB and enclosures free free responses are measured by hanging each from elastic bands. Set-up B consists of an unpopulated PCB directly attached to shaker head expander with 28 M3 bolts evenly spaced around the perimeter as shown in Fig. 3. The bolts gave the PCB a 20 mm stand-off to minimise air-pumping effects as shown in Fig. 4. Seven supposedly identical PCBs existed to allow variability tests later, all measured 250 mm by 250 mm and weighed within ±1 g of 190.5 g. 3.1. Determination of PCB properties test The aforementioned PCB modelling approach was first applied to Set-up A (PCB and enclosure as in Fig. 2). The first step was to experimentally determine the material properties and compare these experimental results with the manufacturer s provided val- 2.5. Current practice: shortcomings The primary shortcoming of the current practice is that it is difficult to find appropriate experimental data in the literature and detailed calculations of the PCB response. With few exceptions, most past analyses have been limited to finding the natural frequencies and very few determine the actual frequency response [20,21]; therefore, it is difficult to estimate the accuracy of the results. The work presented here attempts to overcome this issue by using the aforementioned techniques to model two typical PCBs, leading to an analysis between the predicted response curves and those measured experimentally. The second issue is that the influence of manufacturing and assembly variability has not been assessed. This variability, if severe enough, could have a significant impact on the results, so a series of experiments measuring variability has been carried out Fig. 2. Set-up A, a non-populated PCB integrated into an enclosure attached to the shaker head expander.

90 R.A. Amy et al. / Microelectronics Reliability 50 (2010) 86 97 Fig. 5. Damping of Set-up A measured using different measurement techniques, plotted against maximum acceleration of centre of PCB. Fig. 3. Set-up B, a non-populated PCB attached to the shaker head expander, fixing method is shown in greater detail in Fig. 4. Fig. 4. Detail of fixing method for Set-up B. ues (see Table 1 for a comparison of manufacturers and actual material properties). The manufacturer s provided values were not justified by giving any expected percentage variation, nor was it stated whether these values were maximum or minimum limits. During the experimental tests the following points were made. The material exhibited 15% stiffness variation between the x and y axis, highlighting the significant amount of anisotropy present in the material. It was observed that during both the Youngs modulus and shear tests that the FR4 had relatively high level of hysteresis, and that it was necessary to consider both the loading and unloading cycles. Table 1 Material properties of PCB in Set-up A, experimentally measured values and manufacturers given values. Property Measured Manufacturers values Units Young s modulus (x axis) 4:02 10 10 3:50 10 10 N=m 2 Young s modulus (y axis) 4:56 10 10 3:50 10 10 N=m 2 Density 2481 2400 kg=m 3 Shear modulus 1:21 10 10 Not given N=m 2 Poisson ratio 0.3 0.3 Thickness 1.54 1.66 1.65 mm The thickness of the PCB was measured at multiple points over the specimen; surprisingly it was found that the thickness varied over the material by 0.12 mm, if this amount of variation was not considered throughout the test it would drastically affect the accuracy of the results. The density of this particular sample measured to be 2480 kg=m 3 which was 3% higher than the value of 2400 kg=m 3 given by the manufacturers. Overall, it was noted that the experimentally measured properties of the PCB in Set-up A were appreciably different from the manufacturers published values (see Table 1). Using the damping measurement techniques introduced earlier (formulas (2) (8)), the damping was measured for the combined structure. To investigate whether the damping varies with PCB response the measurements were performed over a range of different input accelerations and using different vibration inputs (see Fig. 5). When using the logarithmic decrement method it was found that decay over several cycles should be used and multiple measurements taken and averaged if good results are to be obtained. The whole process of properties measurement was repeated for Set-up B (set-up shown in Fig. 3, as seven identical PCBs existed it was also possible to measure the variation in the properties. Measured values are given in Table 2 and Fig. 6). The manufacturer could not give the material properties of its boards, although it did state the boards were built according to standard IPC-4101B/21. When measuring the damping properties on this second set-up it was found that a much higher responses could be obtained, this allowed damping values to be given up to much higher board accelerations. 4. Finite element modelling 4.1. Set-up B The first PCB to be modelled was Set-up B (see Fig. 3), with the intention of obtaining a plot of the predicted frequency response that could be compared against the real response. The model was Table 2 Average material properties of seven PCBs in Set-up B. Property Measured Units rð%þ Young s modulus (x axis) 2:4 10 10 N=m 2 6.4 Young s modulus (y axis) 2:0 10 10 N=m 2 4.0 Density 1820 kg=m 3 Neg. Shear modulus 4:4 9 10 N=m 2 4.35 Poisson ratio 0.3 Thickness 1.69 mm 0.26

R.A. Amy et al. / Microelectronics Reliability 50 (2010) 86 97 91 Fig. 6. Damping of PCB in Set-up B measured using different measurement techniques, plotted against maximum acceleration of centre of PCB (using same legend as in Fig. 5). Fig. 7. Comparison of response predicted by FE model (dotted line) and real response of Set-up B. Low base acceleration input was used. built in the PATRAN modelling environment and solved using NAS- TRAN. The model of the PCB was simply created using QUAD4 shell elements. Material properties were created and assigned to these elements using the experimentally derived material properties, a 2D Isotropic material was used. The next step after creating the PCB model was to apply the boundary conditions. The PCB translational boundary conditions were assumed to be rigidly grounded because the PCB in the experiment was directly attached to a 30 mm thick Aluminium plate, thus it was assumed that the response of the plate would be negligible compared to that of the PCB. The PCB rotational boundary conditions were modelled by spring elements (CBUSH) attached to the mesh at each fixing location, and the rotational stiffness of the elements increased until the model frequencies matched the experimental frequencies (see Table 3). The tuned model had a rotational spring constant value of 45 Nm=rad. The final step was to apply damping to the model based on the experimentally derived values (see Fig. 6); unfortunately however, several different damping measurement techniques were used and each gave a different value. To overcome this problem all the different predicted values were tried in the model and the correlation compared to see which gave the best results. The logarithmic decrement method gave a damping of 0.5% (derived using Eqs. (2) and (3)) and was found to give the best results, which agrees with conventional theory that this damping measurement method is the most appropriate for such low values of damping [28]. The final predicted response correlated well with the actual response (see Fig. 7 and Table 3). The previous comparison was based on a very low base excitation (0:01 g 2 =Hz flat input spectrum), permitting the assumption of negligible non-linear effects due to the very small board deflections. What would happen if the small deflection assumption could not be made? To answer this question the test was repeated at a much higher vibration level so that the deflections were significantly greater than the board thickness, ensuring that there should be some non-linear effects [23]. The FE analysis was also repeated with updated damping values based on the new acceleration input and damping values from Fig. 6. A comparison of the results (as shown in Fig. 8) shows that the non-linear effects present at the higher excitation levels significantly alter the response near resonance. The final test on this setup examines the effect of components on model accuracy. The board was populated with 74 g of various different small components, these components were a mixture of different SMT and PTH components, none with an edge length greater than 10 mm. This effect was then incorporated into the FE model by artificially raising the density to simulate the effect of the components. Any additional stiffening effects of the components were ignored. The effectiveness of this method can be seen in Fig. 9 and Table 4. 4.2. Set-up A Set-up A was modelled next (see Fig. 2), again with the intention of examining the difference between the predicted and experimentally measured responses. The PCB FE model was built in a similar way to the previous model (see Fig. 10), using QUAD4 elements and 2D Isotropic material. The free free response of this PCB was calculated and compared to the actual free free response (see Table 5 Column A). Additionally, two more predictions of the free free response were made, one prediction was made using the material properties provided by the manufacturer (Column C) and another made using the material properties measured here Table 3 Comparison of experimental and predicted response of Set-up B. Figures in parentheses are percentage error values. Mode f exp (Hz) f model Q exp Q model 1 146.3 146.8 ( 0.31%) 164.3 149.8 (8.83%) 2 547.1 534.3 (2.34%) 75.4 60.7 (19.49%) 3 564.8 572.2 ( 1.31%) 47.2 64.09 ( 35.8%) Fig. 8. Comparison of response predicted by FE model (dotted line) and real response of Set-up B. High base acceleration input was used to investigate nonlinear effects.

92 R.A. Amy et al. / Microelectronics Reliability 50 (2010) 86 97 Fig. 9. Comparison of response predicted by FE model (dotted line) and real response of Set-up B when populated with components. Low base acceleration input was used. Table 4 Comparison of experimental and predicted response of Set-up B when populated with components. Figures in parentheses are percentage error values. Fig. 11. Finite element model of Set-up A enclosure without PCB attached, note central cross-members to provide additional PCB support (cross-beams and internal stiffening ribs were modelled as 1D beam elements but for clarity are displayed here as representative solid elements). Mode f exp (Hz) f model Q exp Q model 1 129.2 129.5 (0.21%) 170.4 171.1 (0.36%) 2 476.2 467.0 ( 1.97%) 73.9 60.9 ( 21.2%) 3 492.9 500.1 (1.42%) 62.5 65.0 (3.82%) Fig. 10. First torsional mode of finite element model of free free PCB from Set-up A. Fig. 12. Finite element model of Set-up A enclosure with PCB attached. Table 5 Table showing difference in natural frequencies between experimental test on free free unpopulated PCB and various FE models. Model A uses experimentally derived material properties and PCB thickness, Model B assumes constant PCB thickness and model C uses the material properties provided by the PCB manufacturer. Mode Experimental frequency (Hz) Percentage discrepancy A B C 1 41.3 1.19 6.94 15.66 2 66.6 0.15 6.39 31.62 3 93.3 0.75 13.64 37.41 4 110 0.20 1.11 22.36 5 119 0.79 2.18 26.46 6 204 3.13 9.05 33.07 7 210 0.82 0.17 23.67 8 243 1.12 10.65 15.17 9 280 0.14 8.48 2.56 and the material thickness provided by the manufacturer (Column B). The relatively poor accuracy of these last two models highlights the risk of using the material property values provided by the manufacturer. The chassis was modelled using a combination of QUAD4 shell elements for the chassis walls and Bar2 beam elements (see Fig. 11). The only mode shape of any importance to the modelling of the PCB response was that of the central cross-beams, these were experimentally measured to have a natural frequency of 210.4 Hz whilst the FE model of the chassis gave frequencies of 209.8 Hz, a difference of less than 0.5%. To predict the response of the PCB in the enclosure the two models must be combined (see Fig. 12). As before, this was achieved using rigid translational elements and tuned rotational spring elements at the location of each fixing. Unlike the previous model, it was not initially possible to easily tune the spring elements to correlate the frequencies. The initial attempts to model gave significant errors in the frequency of second and third modes, whilst also significantly underestimating the response of the later modes. With some effort it was found that the PCB could be modelled to a good accuracy in the following way: first, in addition to tuning the rotational stiffness it was found that modelling the PCB fixings with a translational spring element allowed good frequency correlation, these additional spring elements also required tuning. This suggests that the initial assumption of the PCB fixings being effectively rigid in translation was incorrect for this set-up. Secondly, it was found that the amplitude response prediction could be improved by specifying lower damping for higher frequency modes. This suggests that damping drops off with frequency, as the higher

R.A. Amy et al. / Microelectronics Reliability 50 (2010) 86 97 93 Table 6 Comparison of experimental results and FE models using different boundary conditions. Subscripts ss, ff and tuned refer to simply supported, fully fixed and tuned edge conditions respectively. Italic values in parentheses are the peak transmissibility for that mode, where these were known. Mode Frequency (Hz) Experiment Tuned SS FF 1 196 195.6 190 207 (31.49) (31.17) (31.17) (31.17) 2 322 314 301 342 3 357 365 349 381 4 478 478 460 515 (35.86) (34.15) (11.71) (11.27) 5 637 635 618 678 6 1010 884 942 1066 Fig. 15. Set-up C attached to the shaker head expander. for the second mode. Variability in the response magnitude was as before (see Table 6). 5. Variability experiments The next set of experiments measure the variation present in some typical PCBs. In addition to the Set-ups A and B used in the previous experiments two more set-ups will also be introduced: Fig. 13. Comparison of response predicted by FE model (dotted line) and real response of Set-up A. Set-up C consists of an MS-6323 Micro-ATX motherboard attached to shaker head expander by six M4 machine screws as shown in Fig. 15. In this test the board only had a 6 mm stand-off as this is how the board is usually mounted. Seven of these boards existed to allow the differenced between each to be measured, each measures 244 mm by 205 mm and weighs within ±1 g of 465 g. In some tests the motherboards free free responses were measured, this was achieved by hanging each board from elastic bands. Set-up D consists of a graphics card (Vanta TNT2M64) suspended by elastic bands to measure free free response (shown in Fig. 16), ten cards were tested this way, each weighs within ±0.5 g of 69 g and measures 150 mm by 83 mm and are 1.6 mm thick. 5.1. Boundary condition variability test Fig. 14. Comparison of experimental results of Set-up A with two different FE models, the Simply supported model assumes the PCB to chassis connection has zero rotational stiffness, whilst the Fully fixed model assumes a rigid rotational connection. modes have lower displacement this would agree with the results of the earlier damping tests. The correlation of the final tuned model can be seen in Table 6 and Fig. 13. Finally, two additional models were made to investigate two commonly made assumptions, namely, that the PCB to chassis connection is either rotationally fully rigid or flexible. Fig. 14 shows the comparison of two models, one using rotationally rigid elements and one using elements that do not constrain rotation. A frequency variation of up to 5% was seen for the first mode and 10% This first test illustrates the sensitivity of a PCBs response to small changes in boundary conditions. This involved removing and re-installing the same Set-up C (as shown in Fig. 15) several times and testing the vibration response after each installation. Any slight variations in the boundary conditions between tests would be apparent as the vibration response would also change between tests (see Fig. 17 and Table 7). The boards were subjected to a 0.5 g vibration input swept from 20 to 400 Hz at two Octaves per minute; this frequency was chosen as it contained the most significant modes, whilst the value of 0.5 g vibration input was a compromise between a good signal to noise ratio and possible damage to the boards. The first test involved removing and reinstalling Set-up C three times, the bolts were tightened to the same torque each time. The second of tests was the same except that not only was the torque identical but also the tightening pattern. Consequently it was found that only by tightening the screws in exactly the same order and to exactly the same torque (1.5 N/m) could any repeatability be achieved (see Table 7), thus all experiments in this work use exactly the same tightening pattern and torque.

94 R.A. Amy et al. / Microelectronics Reliability 50 (2010) 86 97 Fig. 16. Set-up D for measuring free free response of graphic cards. Table 8 Statistical parameters of the vibration response of different set-ups after three successive re-installation attempts. Mode f (Hz) r f r f ð%þ Q r Q r Q ð%þ Set-up C 1 94.5 2.57 2.72 8.50 0.64 7.90 (random pattern) 2 130.3 1.40 1.08 14.76 4.34 29.40 Set-up C 1 97.8 0.20 0.20 8.91 0.62 6.98 2 130.7 0.58 0.44 16.75 1.09 6.51 Set-up A 1 194.4 1.35 0.70 21.14 0.70 3.30 2 333.0 1.17 0.35 4.80 0.15 3.16 3 361.9 0.65 0.18 3.67 0.26 7.06 4 483.5 1.28 0.26 48.14 2.25 4.68 Set-up B 1 145.3 1.72 1.19 147.3 6.67 4.53 2 543.5 6.20 1.14 69.78 9.19 13.18 Fig. 17. Frequency response of Set-up C over successive removal and re-installation attempts, to show sensitivity of board to small variations in boundary conditions (see Table 7). Table 7 Frequencies and peak transmissibilities of first two modes for Set-up C, between each attempt the PCB was removed and then re-installed with bolts re-tightened to the same torque. The first three attempts used a random bolt tightening pattern; the last three attempts used exactly the same bolt tightening pattern. 5.2. Manufacturing variability test The second set of tests involved testing and comparing the response of seven supposedly identical motherboards as in Set-up C, these were tested in a mounted condition similar to the previous tests. It was immediately apparent that identical manufacture did not mean identical response (see Fig. 18 and statistics of variation Attempt Mode 1 Mode 2 f (Hz) Q f (Hz) Q 1 92.6 9.20 128.8 15.25 2 93.4 7.92 131.6 18.84 3 97.4 8.4 130.0 10.20 4 97.6 8.40 130.0 16.0 5 98.0 8.72 131.0 16.25 6 97.8 9.60 131.0 18.0 It was suspected that different test set-ups might be more sensitive to slight changes in boundary conditions than others. To investigate this assumption the tests were repeated for two Setups A and B (the statistics of variation are shown in Table 8). Preliminary testing showed the sensitivity to bolt tightening pattern was found to be insignificant in both these tests, so this part of the original test was not repeated. Fig. 18. Frequency response of seven supposedly identical motherboards (Set-up C), each board was installed in exactly the same manner, with identical torques and bolt tightening pattern (see Table 9 table for specific information).

R.A. Amy et al. / Microelectronics Reliability 50 (2010) 86 97 95 Table 9 Statistics of variation for seven identical Micro-ATX motherboards in Set-up C, each with identical installation procedures. Mode f (Hz) r f r f ð%þ Q r Q r Q ð%þ 1 94.74 2.94 3.10 11.34 2.54 22.42 2 129.31 1.45 1.12 14.80 0.82 5.56 Fig. 20. Graph to show decrease in first resonant frequency of a Set-up B due to damage from intense vibrations. Arrow points in the direction of increasing levels of vibration. Fig. 19. Thickness against first natural frequency for identical Micro-ATX PCBs as in Set-up C. in Table 9). Additionally, some correlation was noted between the motherboard thickness and first natural frequency (see Fig. 19), leading to the conclusion that small differences in motherboard thickness were partly responsible for the variation in the response. The tests were repeated on Set-ups B and D to investigate whether this level of variation was typical (set-ups shown in Figs. 15 and 16 but free free). To remove any possible boundary condition effects this second set of tests was performed using free free conditions (results are shown in Table 10). 5.3. Variations in response after high acceleration A test was carried out to show how the response of a PCB might be irreversibly altered due to damage from high vibration loading. Set-up B was attached to a shaker head expander and excited at Table 10 Statistics of modes of several identical PCBs for free free conditions. Seven boards were tested in Set-up B and C, whilst ten were tested in Set-up D. Set-up Mode f (Hz) r f r f ð%þ Set-up C 1 53.51 1.50 2.80 2 76.00 0.86 1.13 3 103.10 1.57 1.52 4 134.03 1.68 1.25 5 157.1 4.06 2.58 Set-up B 1 23.5 0.54 2.31 2 48.0 1.41 2.94 3 62.1 0.62 1.00 4 72.0 1.33 1.84 5 75.2 0.75 1.00 6 133.0 0.79 0.59 7 149.2 0.82 0.55 8 157.5 0.70 0.49 9 179.1 0.89 0.50 10 230 1.57 0.68 Set-up D 1 202.4 3.22 0.79 2 262.1 3.74 0.71 several levels of vibration from 0.2 g to 40 g sinusoidal input, the vibration input was a sinusoidal input at just below the first resonant frequency of the board, each level lasted ten minutes. A visual inspection of the PCB after each test did not see signs of obvious damage. The response of the board was then measured between each stress test using a low-level sine sweep at 0.2 g. It was apparent that the response of the PCB was affected when more intense vibrations were applied (see Fig. 20), as the first resonant frequency of the PCB fell 3.2% and the acceleration response fell 11%. This first test used a set-up that incorporated nylon washers that were suspected of being susceptible to yielding, so the test was repeated with another set-up that had no washers. Over this second test the frequency dropped by only 1.5% while the response dropped by a significant 18.6%. It was observed that the majority of the reduction for the second test occurred when the board centre response reached over 160 g peak acceleration, whilst for the first test the reduction and therefore damage occurred evenly throughout the test. In summary, the response was irreversibly altered due to high accelerations even though no visible damage occurred. This damage could take many possible forms: delamination of the glass fibre layers within the PCB, failure of the epoxy to fibre bond, local yielding of the PCB around the fixing that resulted in lower bolt tension. Although the exact cause of failure can not be determined it is enough to know that at high responses such variations can occur and should be measured if possible, if measurement is not possible then appropriate safety factors should be included. Also, it should be noted that the damage did not affect any of the electrical connections on the PCB in this case study, but for multi-layer PCBs the possibility of damage to electrical connections could be an issue. 6. Discussion Say the response of a piece of equipment is known, to what extent should responses of other identical pieces of equipment expected to be the same? First, let s consider a worst-case scenario based on the results reported here. For example, if the assembly of a piece of equipment is not tightly controlled, resulting in different bolt torques and tightening patterns, the response magnitude may have a standard deviation of up to 30% (based on the highest variation seen in Table 8). Low manufacturing tolerances may contribute up to another 22% standard deviation (based on the highest variation seen in Table 9). These percentages are then corrected for the small sample size by using a Student s t-distribution, at a 99% confidence level

96 R.A. Amy et al. / Microelectronics Reliability 50 (2010) 86 97 this distribution requires that the percentages be multiplied by 3.14 to safely account for any possible variability. Based on these values, for a worst-case scenario, the actual response may be 184% higher or lower than the measured value. Furthermore, this does not consider that the frequencies may vary by up to 18% higher or lower than the measured value, possibly placing the resonant frequencies of the structure into a region of higher input acceleration. Finally, if large accelerations are present damage may occur to the structure that could alter the response further. Now let s consider a best-case scenario. For equipment that is assembled in a very controlled environment, the response magnitude may vary by as little as 3% (based on the lowest variation seen in Table 8), whilst the lowest deviation observed due to manufacturing variations was 3% (based on the lowest first mode variation seen in Table 10). At three sigma this amounts to the response being possibly 18% higher or lower than the measured response. Thus, to answer the original question of how similar responses may be, is not so straightforward, as it can be seen that the variability depends on the specific case at hand. For example, the motherboards in Set-up C seemed to be more sensitive to small variations in boundary conditions, possibly due to their relatively small number of fixing points. Therefore, any estimate of variability should be based on set-ups and assembly processes as similar as possible to the equipment to which they will be applied. Now let s consider the question that was posed earlier: With what accuracy can the vibration response reasonably be predicted by an idealised theoretical PCB FE model? First, if a specimen of the PCB material is available for bend testing, then the PCB material properties (stiffness, density, and thickness) can be measured and specified accurately. With this information it is then possible to get very good correlation between the predicted and actual PCB free free response; thus removing one significant source of error. Next, to overcome the problem of measuring damping, this can be measured from an already existing piece of equipment. This technique is only valid if the existing equipment is sufficiently similar to the proposed design, i.e. similar board size, fixtures and chassis. Fortunately, most companies only use a few different form factors, so it is very likely that similar pieces of equipment already exist. With these two pieces of information (PCB properties and damping) a full FE model of the combined PCB/enclosure system can be created; with this model it is now possible to attempt to predict the edge rotational stiffness. This can be achieved using the same technique as was just described for damping: using estimates based upon existing equipment with similar fixings. All these measured values can be saved in a database and then be re-used for any future modelling situations, without the need to be measured every time. As it is assumed that most manufacturers use only a few different types of PCBs, form factors, fitting mechanisms and chassis types, it is possible that with a relatively modest amount of effort, a database of all the different input parameters could be created, thus allowing future models to be created quickly and effectively. In this paper, the FE modelling has been demonstrated for two different examples of electronic equipment. In one example it was possible to predict the edge rotational stiffness and the resultant response prediction was very good. In the second example the boundary conditions were more complicated and required a more complicated modelling approach, this approach had to include both translational spring constants at the PCB boundary conditions and also frequency dependant damping. Finally, at high levels of excitation, non-linear effects may become significant and further decrease accuracy. Fortunately, the non-linear effects seem to only reduce the response, so ignoring them is a conservative assumption. Again, the answer to the posed question is not straightforward. If a structure is submitted to extensive preliminary testing and has relatively simple PCB boundary conditions that permit trial and error calculation, then it is possible to create a relatively accurate model. However, in a real situation, it is unlikely that such extensive testing will be practical and even if such testing is undertaken, there remains a possibility that the boundary conditions and damping are relatively complicated to specify. In addition, the following points have been noted during the measurement process: accurate damping measurement is vital if the FE model is to accurately predict response magnitude. Measurements should be made at different power input levels to reflect how damping varies with increased deflections. The Logarithmic method is the most suitable method for doing this, especially when the damping is low, although it is recommended to average multiple readings when using this method. In terms of determining the PCB material properties the following points have been noted. The laminate material is noted to have strongly anisotropic properties and an accurate model cannot be created without using the stiffness moduli in both the x and y axis. The importance of measuring and including shear stiffness in the model should not be overlooked. Thickness variations can significantly affect the accuracy of results if not considered. 7. Conclusion The aim of this work was to assess the accuracy of typical FE models used to predict the response of PCBs to harsh random vibration environment. It has been shown that if good data on the board properties exists, then even simple PCB FE models can deliver very accurate response prediction. In practice, however, various input parameters are difficult to obtain: namely, reliable data concerning boundary conditions, stiffness or damping. It has been shown that with enough preliminary testing it is possible to determine the stiffness and damping values, whilst the boundary conditions may be calculated by a tuning approach. Unfortunately, this does create the requirement that an experimental set-up similar the one being modelled already exists. Ultimately, with the approach given here, an estimate of the expected accuracy of a PCB FE model can be made, from this estimate the confidence in future models can be known. In addition, tests were performed to show that manufacturing and assembly variability can produce significant differences in the response of supposedly identical PCB assemblies, highlighting the importance in measuring this variability. Acknowledgements The author would like to warmly thank Surrey Satellite Technology Limited (SSTL) for their vital contribution to this research. Special thanks also go to Ian Black for kindly providing free equipment and technical support. References [1] Dehbi A, Ousten Y, Danto Y, Wondrak W. Vibration lifetime modelling of PCB assemblies using Steinberg model. Microelectron Reliab 2005;45(9 11):1658 61. [2] Gu J, Barker D, Pecht M. Prognostics implementation of electronics under vibration loading. Microelectron Reliab 2007;47(12):1849 56. [3] Chen Y, Wang C, Yang Y. Combining vibration test with finite element analysis for the fatigue life estimation of PBGA components. Microelectron Reliab 2008;48(4):638 44. [4] en Luan J, Tee TY, Pek E, Lim CT, Zhong Z, Zhou J. Advanced numerical and experimental techniques for analysis of dynamic responses and solder joint reliability during drop impact. IEEE Trans Compon Packag Technol 2006;29(3):449 56.

R.A. Amy et al. / Microelectronics Reliability 50 (2010) 86 97 97 [5] Shetty S, Lehtinen V, Dasgupta A, Halkola V, Reinikainen T. Fatigue of chip scale package interconnects due to cyclic bending. ASME J Electron Packag 2001;123(3):302 8. [6] Shetty S, Reinikainen T. Three- and four-point bend testing for electronic packages. ASME J Electron Packag 2003;125(4):556 61. [7] Li R. A methodology for fatigue prediction of electronic components under random vibration load. ASME J Electron Packag 2001;123(4):394 400. [8] Jih E, Jung W. Vibrational fatigue of surface mount solder joints, ITherm 98. In: Sixth intersociety conference on thermal and thermomechanical phenomena in electronic systems (Cat. No. 98CH36208); 1998. p. 246 50. [9] Kotlowitz R. Comparative compliance of representative lead designs for surface-mounted components. IEEE Trans Compon Hybrids Manuf Technol 1989;12(4):431 48. [10] Kotlowitz R. Compliance metrics for surface mount component lead design, 1990. In: Proceedings of 40th electronic components and technology conference (Cat. No. 90CH2893-6); 1990. p. 1054 63. [11] Kotlowitz R. Taylor L. Compliance metrics for the inclined Gull-wing, Spider J- bend, and Spider Gull-wing lead designs for surface mount components. In: Proceedings of 41st electronic components and technology conference (Cat. No. 91CH2989-2); 1991 p. 299 312. [12] Pitarresi J, Caletka D, Caldwell R, Smith D. The smeared property technique for the FE vibration analysis of printed circuit cards. ASME J Electron Packag 1991;113:250 7. [13] Pitarresi J, Primavera A. Comparison of vibration modeling techniques for printed circuit cards. ASME J Electron Packag 1991;114:378 83. [14] Lall P, Panchagade D, Choudhary P, Gupte S, Suhling J. Failure-envelope approach to modeling shock and vibration survivability of electronic and MEMs packaging. IEEE Trans Compon Packag Technol 2008;31(1):104 13. [15] Wu J, Zhang RR, Radons S, Long X, Stevens K. Vibration analysis of medical devices with calibrated FEA model. Comp Struct 2002;80:1081 6. [16] Pitarresi J, Roggeman B, Chaparala S, Geng P. Mechanical shock testing and modeling of PC motherboards. In: Proceedings of 54th electronic components and technology conference (IEEE Cat. No. 04CH37546), vol. 1; 2004. p. 1047 54. [17] Barker DB, Chen Y. Modeling the vibration restraints of wedge-lock card guides. ASME J Electron Packag 1993;115(2):189 94. [18] Lim G, Ong J, Penny J. Effect of edge and internal point support of a printed circuit board under vibration. ASME J Electron Packag 1999;121(2):122 6. [19] Guo Q, Zhao M. Fatigue of SMT solder joint including torsional curvature and chip location optimization. Int J Adv Manuf Technol 2005;26(7 8):887 95. [20] Aglietti GS. A lighter enclosure for electronics for space applications. Proc Inst Mech Eng Part G 2002;216:131 42. [21] Aglietti GS, Schwingshackl C. Analysis of enclosures and anti-vibration devices for electronic equipment for space applications. In: Proceedings of the 6th international conference on dynamics and control of systems and structures in space; 2004. p. 261 70. [22] Yang Q, Wang Z, Lim G, Pang J, Yap F, Lin R. Reliability of PBGA assemblies under out-of-plane vibration excitations. IEEE Trans Compon Packag Technol 2002;25(2):293 300. [23] Cifuentes AO. Estimating the dynamic behavior of printed circuit boards. IEEE Trans Compon, Packag, Manuf Technol Part B: Adv Packag 1994;17(1):69 75. [24] Amy RA, Aglietti GS, Richardson G. Sensitivity analysis of simplified PCB FE models. In: Proceedings of the 1st CEAS European air and space conference CEAS-2007-466; 2007. [25] Sarafin TP. Spacecraft structures and mechanisms: from concept to launch. Kluwer Academic Publishers; 1995. [26] en Luan J, Tee TY, Pek E, Lim CT, Zhong Z. Dynamic responses and solder joint reliability under board level drop test. Microelectron Reliab 2006;47(2 3):450 60. [27] Wang H, Zhao M, Guo Q. Vibration fatigue experiments of SMT solder joint. Microelectron Reliab 2004;44(7):1143 56. [28] de Silva CW. Vibration fundamentals and practice. CRC; 1999. [29] Steinberg D. Vibration analysis for electronic equipment. John Wiley & Sons; 2000. [30] Liguore S, Followell D. Vibration fatigue of surface mount technology (SMT) solder joints. In: Proceedings of the annual reliability and maintainability symposium (Cat. No. 95CH35743); 1995. p. 18 26. [31] Sidharth K, Barker DB. Vibration induced fatigue life estimation of corner leads of peripheral leaded components. ASME J Electron Packag 1996;118(4):244 9. [32] Oguibe C, Williams D. Reliability assessment of a digital electronic board assembly using the physics-of-failure approach: a case study. Proc Inst Mech Eng, Part B (J Eng Manuf) 1999;213(B8):833 9.