You are on page 1of 11

Bounding Natural Frequencies in Structures I: Gross Geometry, Material and Boundary Conditions

Keith B. Smith Research Engineer Southwest Research Institute 6220 Culebra Road San Antonio, TX 78238 kbsmith@swri.org

William C. Shust, P.E. Mechanical Engineer Objective Engineers, Inc. 363 Venturi Drive Pueblo West, CO 81007 objengrs@aol.com

ABSTRACT A classic aspect of good structural design lies in optimizing stiffness-to-mass ratio through material and shape choices; and to a new engineer, it may appear that natural frequency can be manipulated as an afterthought. For conventional structures, however, the roles of material, cross section and boundary conditions are all comparatively small (when compared to gross geometric aspect ratio) in altering fundamental frequencies. Despite apparent differences, the most damaging flexible modes for structures ranging in size from a supertanker down to a sedan occur at similar frequencies less than 30 Hz. It is not until an article shrinks to the realm of small electronic components that modes driving significant primary-structure damage rise to kilohertz ranges. Ground vehicle, aircraft, marine and civil structures are examined to show their broad physical design space, yet relatively narrow bounds of first natural frequencies. Simple beam theory is used to explain this apparent contradiction and therefore limits upon designed-in (or designed-out) natural frequencies. The implication for designers is that damping is critical, as resonant mode shapes are difficult to geometrically shift out of structurallydamaging bandwidths. The implication for test engineers is that high sampling rates, while useful for acoustic and impact problems, may be unnecessary for questions of structural fatigue. NOMENCLATURE
Asec
= beam moment of intertia around bending axis = beam coefficient for a given boundary condition = beam coefficient for a given cross section = coefficient for given boundary conditions = coefficient for given beam cross section = cross section dimension (e.g., average height or diameter) = modulus of elasticity for beam material = cyclic natural frequency [e.g., Hz] = modulus of rigidity for beam material = moment of intertia around bending axis of beam = system stiffness [e.g., lb/in] = beam stiffness as a function of end/load constraints = length of beam (perpendicular to vibration direction) = system mass [e.g., lbs2/in]

Bend Bsec
Cend = Bend Rend

Csec = Bsec

D E mat
fn

Gmat I zz k
K end

Lbeam

m=w g

mbeam
meff = Rend mbeam

= total mass of beam (no supplementary mass) = effective or equivalent lumped mass = mass density of beam material = Rayleigh coefficient for given boundary condition = kinetic energy = velocity at a point on beam = circular frequency for nth-mode [e.g., rad/sec] = distance along beam = equation of elastic curve for given boundary condition

mat
Rend T v

n
x

y ( ymax , x)

INTRODUCTION & HYPOTHESIS Outside the study of transient loadings, structural dynamicists are most often interested in random input vibrations at frequencies well below 100 Hz. The most damaging resonant modes of some larger structures (e.g., buildings, bridges and ships) are typically excited at less than 10 Hz; while collecting durability data, aircraft and vehicle test engineers nominally use 30-50 Hz low-pass anti-aliasing filters to allow for more efficient digital sampling rates. This paper will address several hypothetical questions:

What are some basic governing principles that prevent stiffness from increasing faster than mass as small structures the size of automobiles grow into larger structures the size of ships? What drives human-occupied structures constructed with a wide array of materials, geometries and expected loads toward similar fundamental frequencies? What is the practical result for engineers, and why are the higher vibration modes often ignored?

A structural engineer in any of these fields (aeronautical, mechanical or civil) may not have time to become fluent in the language of vibration as s/he studies the complexities of many possible elastic, plastic and fracture failure modes. But a minimum passing knowledge of how structures respond to dynamic energy inputs is critical for a true understanding of the environmental loads and specified design criteria. This is especially critical for the growing number of fatigue crack growth experts who will not always have a local strain measurement in the vicinity of a damaged area under study; rather, they will often have only gross structural inertial measurements to work from. The intent of this paper is to provide a new engineer with some practical information regarding the measurement and/or prediction of gross structural vibration modes. FUNDAMENTAL SURVEY: SIZES & BANDWIDTHS OF STRUCTURES Narrowing the Field of Study To constrain this study, several variables were fixed. As a first step, the entire field of structures was reduced to five major classes for consideration, each with three subclasses based on gross geometric size. A literature search1 was then conducted to gather natural frequency information for each class; the scope was limited to the first few gross bending and/or torsional modes of free vibration, as experience has shown that those are usually the most immediately damaging to the primary structure (i.e., higher mode shapes in the primary structure are often accompanied by smaller local deflections, depending upon input energy at the corresponding band). Aspect ratio information was also gathered, but limited to the main longitudinal structure (i.e., wings/flutter was not included for aircraft and piers/towers were not included in bridge size characterizations), and includes only the general envelope apparent to the casual observer that encompasses the gross structure (e.g., entire body of an automobile, rather than just the frame). For relative comparisons, general frequency ranges of interest for some substructure vibrations have been added to the bottom of Table 1.

A complete (though contracted to save space) list of sources consulted during the development of this work is offered at the end of the paper in the last section, entitled "References & Bibliography."

Table 1. Survey of Structural Sizes and Lower Mode Natural Frequencies


Structural Class Ground Vehicles Commercial Aircraft Boats / Vessels Rectangular Buildings Bridges Subclass Automobile (SUV) Tractor Trailer (Connex) Railcar (Autorack) Light Piston Business Jet/Turbo-Prop Long Range (Cargo) Icebreaker/Coaster Passenger/Freighter Supertanker House (1-3 Stories) Building (~20 Stories) High Rise (>30 Stories) Small (Wooden Foot) Medium (Beam Girder) Large (Suspension) Mechanical Components Electronic Components Mechanically-Induced Acoustic Noise Width/Height/Length (ft) [Volumetric Aspect Ratio]2 5 / 4 / 17 [3.77] 8 / 9 / 40 [4.71] 10 / 18 / 89 [6.36] 4 / 3 / 21 [6.00] 6 / 5 / 58 [10.5] 20 / 18 / 240 [12.6] 49 / 30 / 215 [5.44] 55 / 30 / 371 [8.73] 70 / 53 / 882 [14.3] 56 / 28 / 35 [0.83] 98 / 50 / 184 [2.49] 210 / 110 / 380+ [>2.38] 28 / 5 / 166 [10.1] 34 / 11 / 477 [21.2] 90 / 27 / 3900 [66.7] various (~Structure Size/10) various (W/L~1, D~1/16th in) n/a Typical Bandwidth3 for 1st Few Modes (Hz) 17 - 27 3.6 - 11 0.5 - 9.0 7.9 - 25 4.0 - 20 5.0 - 20 4.1 - 7.5 1.3 - 3.0 0.1 - 4.8 0.8 - 3.3 0.3 - 1.5 0.4 - 2.4 2.4 - 10.3 0.5 - 5.7 0.2 - 1.4 30 - 480 150 - 2500 1000 - 20000

Substructure

Bending and torsion of gross primary structure appears well rooted in the double order of magnitude spanning 0.15-30 Hz. This seems true regardless of whether a civil, offshore, mechanical or aerospace engineering outfit designed the structure. It is not until the size of the object shrinks to that of an electronic component that response in the kilohertz range of frequencies become structurally important. Furthermore, general design trends are such that all of these structures seem to grow longer faster than they grow wider or taller in profile; this is generally the case both within structural classes and between them, and is demonstrated in the table as the aspect ratio between the length and profile increases for the relatively larger structures. Trends in Longitudinal Aspect Ratio As demonstrated in general case of Figure 1, structural stiffness in bending is an extremely small fraction of what it is in axial loading, and the strains resulting from an applied bending force can be orders of magnitude greater than if the same load had been applied axially. It therefore seems counterintuitive that as larger structures are designed, their volumetric aspect ratios increase especially when many critical inputs to these structures are perpendicular to their length (i.e., increasing the relative length can increase structural vulnerability to flexure driven by primary service inputs, such as wind on a building or ocean swell on a ship)!

Figure 1. Structural Vulnerability to Bending Loads [2]


Volumetric aspect ratio is defined here as the length divided by nominal cross sectional girth (e.g., 2L/(W+H)) quoted for each type of structure; the variable "D" is later used to quantify an average cross section dimension. 3 Frequencies reported are representative of the first few modes of free vibration only (usually gross structural bending and torsion); fatigue-damage in primary structure is typically driven by cycle rates < 30 Hz.
2

There are a variety of motivations for designing structures as long as cross-sectional material strength will permit. In the case of civil structures, the cost of real estate in urban areas undoubtedly tips the cost-benefit analysis toward vertical growth for buildings, while material and construction costs may drive it for bridges. And certainly drag coefficients have a strong effect on the shape factors chosen for moving vehicles whether on land, in the air or at sea. Given that cost effectiveness works against structural strength in this manner, gross bending and torsion computations often drive macroscopic aspect ratio decisions very early in the design process. Parametric Beam Model Development Because the general engineering structures surveyed here have an almost infinite array of differences in their structural details, materials and boundary conditions, a simple box beam model was developed for a cursory examination of natural frequency sensitivity to the following four parameters:

End supports (used to approximate the effect of structural boundary conditions) Cross section (used to approximate the effect of structural section modulii) Material (homogeneous/isotropic approximation of complex structures) Size (used here in an absolute sense; gross length, width and height)

Insight into the complex behavior of real structures can be gained from this simple model. Structural damping was neglected; though this can have an important effect on response amplitudes, its influence on natural frequency is typically very small (e.g., a few percent). Of course when attention is turned toward comparing physical damage due to specific vibration modes, damping cannot be ignored. However, the focus here is on response frequency content rather than amplitude content (i.e., input vibration bandwidths that drive resonant modes, as opposed to any specific damage quantification). As a final simplification, model parameters were varied under free vibration assumptions for a uniform undamped beam. Note that for each given set of conditions, the fundamental natural frequencies will drop and the maximum stresses will increase with the addition of external mass. This, however, would not appreciably affect the general trends related to gross structural size. FREQUENCY SIMILITUDE: A SIMPLE STRUCTURAL MODEL Beam Bending Parameterization The development of a parametric formula for the natural frequencies of pure bending in a beam is fairly straightforward. Numerous textbooks have applied Newton's second law of motion to the simple harmonic motion of an undamped single degree of freedom system to demonstrate the classic relationship of Equation 1.

n = 2f n = k m

(1)

Lord Rayleigh modified this basic relationship to arrive at a better estimate of fundamental frequency by resolving the effect of the distributed mass of the spring itself, in addition to that of the mass in the simple spring-mass model. In the case of a beam, by assuming its distributed mass is lumped at the position of maximum displacement, maximum kinetic energy can be calculated by integrating Equation 2 along the length of the beam, and then solving for the effective mass of the beam at that point.

T = meff v 2
2

1m 2 1 & dx = (Rend mbeam ) y & max 2 Tmax = beam y 2 l 2 0

(2)

where:

meff = Rend mbeam = effective or equivalent lumped mass

The Rayleigh coefficient is a measure of the percentage of the total beam mass that is effective (or would be, at the center of gravity of a massless beam) in sustaining the resonant mode shape. This is conceptually demonstrated in the schematic of Figure 2 below. For a simple beam in its first bending mode, Rend is

approximately 0.384 with fixed ends, 0.486 simply supported and 0.775 cantilevered. As expected, the effective mass rises as the wavelength of the mode shape is increased.

mbeam

meff

Figure 2. Rayleigh Effective Mass Concept From general beam theory, sectional stiffness can be estimated as a function of the beam length, area moment of inertia and material modulus of elasticity. This member stiffness can then be coupled with the Rayleigh effective mass for the given loading and end conditions to resolve a natural frequency parameter that is a function of the boundary conditions. This process is illustrated for the fundamental beam bending frequency in Equation 3.

f1B =

1 2

K end 1 = meff 2

Bend Emat I zz C = end 3 2 Rend mbeam Lbeam

Emat I zz 3 mbeam Lbeam

(3)

where:

Cend = Bend Rend = coefficient for given boundary conditions

The parameterization can be taken a step closer to the denominators of interest with simple substitutions for the mass moment of inertia and total mass of the beam. The radius of gyration is not extracted, as it is a function of the cross-sectional size; the intent here is to resolve a dimensionless parameter that is a function only of the relative cross-sectional shape, not the absolute magnitude of its dimensions.

f 1B

C = end 2

E mat Bsec D 4

mat Asec Lbeam 4

C end C sec 2

E mat D 2 mat L beam

(4)

where:

Csec = Bsec = coefficient for given beam cross section

For conceptual convenience, this equation is further reduced to a final form in Equation 5. Here, it is once again expressed in terms of circular frequency for simplicity, and it is generalized to include free vibration modes for both bending and torsion. Note that there are two coefficients, representing end conditions and shape effects, multiplied by two functions, representing material and overall size effects.

n = C end C sec Fmat Fsize


where:

(5)

Fmat =

E mat

mat

and

Fsize =

D Lbeam
2

for bending

(material & geometry functions)

additionally:

Fmat =

Gmat

mat

and

Fsize =

1 Lbeam

for torsion

For a given set of boundary conditions and cross section, the natural frequency of any beam is now a function of its material and gross geometry. The remainder of this paper is a practical exploration of how each of these parameters can affect natural frequencies of free beam vibration. Boundary Condition Sensitivity Cend On a macroscopic scale, the boundary conditions for the first bending and torsional modes of basic structures can be simply approximated. And for a beam under free vibration (no external mass or forcing function), boundary condition coefficients ( Cend ) are shown in Table 2 for the same constraints the Rayleigh coefficients were determined for previously. Structural classes that these coefficients might roughly approximate are also listed. Table 2. Boundary Condition Coefficients for Some First Modes of Gross Structural Vibration
Free Vibration Mode Shape Typical Structural Class Ground Vehicles, Commercial Aircraft Cend (Mode 1) 9.94

Rectangular Buildings

3.52

Boats/Vessels (free-free), Bridges

22.4

Bridges

15.4

Rectangular Buildings (as well as some others above)

1.574

The coefficients for simple first-mode beam bending illustrate an important concept for gross structural vibration. With all other parameters constant (i.e., Csec , Fmat , Fsize ), varying the boundary conditions could nominally vary the first natural bending frequency of a general structure by a factor up to 6.4 (22.4/3.52) at the worst extreme. But because vehicles moving over the ground or through the air/water typically operate under similar end constraints (often approximated as simply supported), there may be little sensitivity of their mode shapes and fundamental frequencies to this parameter. However, end effects alone may separate the fundamental bending frequencies of vehicles by a factor of 2-3 from those of buildings and bridges. Section Shape Sensitivity Csec A non-dimensional cross section representation has also been extracted as an independent coefficient in the determination of natural frequency. In short, the numerical value of this constant is a function of the relative shape of the cross section, and does not change based on the actual size. The nominal cross section of many structures can be encompassed with either a rectangular or circular envelope (e.g., an I-beam can be thought of as a square with two smaller squares removed on each side); the overall aspect ratio and percentage of area remaining inside that envelope is dependent upon a variety of factors, such as whether the primary loads are carried in the shell or through a frame. To understand how the inherent shape of this cross section can influence fundamental frequencies, hollow tubes (both square and cylindrical) with variable wall thicknesses are examined here in first bending. Figure 3 graphically illustrates the cross section parameter ( Csec ) for both square and circular extrusions (e.g., pipes), and may be representative of some simple shell structures.

Recall the material and geometry functions for torsion are different than those of bending; due to this difference, comparisons to bending based on this coefficient alone can be misleading the information is provided here for reference only.

Cross Section Parameter, C sec

0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0.0 0.1 0.2 0.3

Solid Square

Solid Circle

Wall Thickness (height or diameter)/100


0.4 Cylinder (outer diameter) 0.5

Wall Thickness / Section Height


Rectangular Tube (outer height)

Figure 3. Section Shape Coefficients for First-Bending of Tubulars Many common box and tubular sections have wall thicknesses that vary from 3-12% of the overall cross-section height or diameter. The net effect is that most off-the-shelf closed sections will likely have shape coefficients ranging from 0.138-0.239 if square and 0.106-0.183 if round. But for either shape, changing the wall thickness while keeping other parameters constant has the practical effect changing the natural bending frequency by a factor of approximately 1.7 (e.g., moving from a thin- to a thick-walled tube for an application). While it remains to be demonstrated whether real structures are bounded by a similar range (i.e., 330% section modulus increases with a fixed outer footprint that result in only a ~1.7x increase in resonant frequency), perhaps this factor can be compared in a relative sense to the other parameters calculated for simple beams to make inferences about real structures. Additionally, this plot provides insight into radius of gyration effects given two tubular sections of the same wall thickness where the outer diameter of the circle is equal to the outer height of the square; the square section will resonate at frequencies approximately 30% higher than those of the circular. Furthermore, filling in a cross section with material may only increase the frequency of first bending by a factor of 2.7 (the practical difference between a thin-walled tube and a solid rectangle of the same size)! Material Function Sensitivity Fmat The quantification of material effects on the fundamental frequency of beam bending is easily performed with a calculation of the longitudinal velocity of sound through each material ( Fmat = Emat mat = Vsound ). Data of this nature for a range of common materials often used in structure are listed in Table 3. Table 3. Material Effects on First Modes of Gross Structural Vibration
Material Class Metals Material Steel Titanium Aluminum Zirconia-Toughened Alumina Silicon Nitride Yttria-Stabilized Zirconia Polyester (thermoset) Nylon Polyamide Silicone Rubber Polymer-Matrix FiberglassReinforced Composite Young's Modulus (x106 psi) 29.0 15.5 10.1 50.0 49.3 30.5 0.469 0.392 1.10 4.71 Mass Density (lbm/ft3) 491 282 169 262 203 381 78.0 71.2 84.3 122 Fmat (x103 ft/s) 16.6 16.0 16.7 29.7 33.6 19.3 5.28 5.05 7.79 13.4

Ceramics Polymers / Elastomers Composites

This straightforward calculation is very powerful in demonstrating the potential effect of a material change only on a structural response. In addition to the consistency within classes of material, the data in Table 3 suggest that given a fixed geometry, changing a structure from one metal to another will not have much effect on the natural frequencies of that structure (though there is a bit more variability within the ceramic and polymer/elastomer classes). Furthermore, changing material in that same structure to a polymer-fiberglass composite may only shift resonances by a factor of 1.2. To get a shift of ~6-7x out of material changes alone, a component formed of silicon nitride would have to be swapped out with one made of Nylon a very unlikely replacement indeed! This is representative of the strongest effect that material choices will typically have on the resulting natural frequencies in general structural applications. Size Function Sensitivity Fsize As it turns out, the size parameter is the only one of the four structural parameters identified that has the potential to shift the fundamental frequencies by several orders of magnitude. This parameter ( Fsize = D Lbeam 2 ) is expressed here as the diameter or height of the cross section divided by the square of the gross length. Outside dimensions only were considered for this survey; mass coupling during vibration (i.e., the portions of substructure participating in a particular mode shape) is important to consider this is discussed further later. One of the interesting things about this particular relationship is that for a fixed volumetric aspect ratio (as well as materials and boundary conditions), the overall size of the structure has to decrease by two orders of magnitude to increase the fundamental frequency by two orders of magnitude. Therefore, the size of the sport utility vehicle in Table 1 may have to decrease to approximately the size of a hand-held toy car (0.6 x 0.5 x 2 in) to move a 20 Hz resonant mode to 2000 Hz, and therefore into the realm of electronic components. Conversely, this inversely proportional relationship suggests that it may need to be enlarged to the approximate size of a 200-ton supply boat (50 x 40 x 170 ft) to shift that same mode down to 2 Hz! The geometry parameter has replaced the volumetric aspect ratio for the structural classes of Table 1 in Table 4 below. Table 4. Size Effects for Primary Structural Classes
Structural Class Ground Vehicles Commercial Aircraft Boats / Vessels Rectangular Buildings Bridges Subclass Automobile (SUV) Tractor Trailer (Connex) Railcar (Autorack) Light Piston Business Jet/Turbo-Prop Long Range (Cargo) Icebreaker/Coaster Passenger/Freighter Supertanker House (1-3 Stories) Building (~20 Stories) High Rise (>30 Stories) Small (Wooden Foot) Medium (Beam Girder) Large (Suspension) Fsize = D/Lbeam2 (x10-3 ft-1) 15.6 5.31 1.77 7.94 1.64 0.330 0.855 0.309 0.0791 34.3 2.19 1.11 0.599 0.0989 0.00385 Typical fn Range for 1st Few Modes (Hz) 17 - 27 3.6 - 11 0.5 - 9.0 7.9 - 25 4.0 - 20 5.0 - 20 4.1 - 7.5 1.3 - 3.0 0.1 - 4.8 0.8 - 3.3 0.3 - 1.5 0.4 - 2.4 2.4 - 10.3 0.5 - 5.7 0.2 - 1.4

In reviewing trends due to size within this general data, the three-story house and suspension bridge are treated as outliers. Recall the house has an aspect ratio approximately 3x higher than any others under consideration (0.8), and the bridge length was more than 10x longer than any other structure in the table (3900 ft). The size parameters for the remaining structures are within a factor of 200 of one another. This corresponds well with their approximate typical response bandwidth of 0.15-30 Hz, which also spans a factor of 200. And within each subclass of structures (excluding the house and suspension bridge), there is less than a factor of 25 difference between extreme sizes. It is important to note that the size parameters tabled here are based on a gross geometric envelopes, not necessarily the size of the primary structure actively participating in the bending mode of interest (i.e., for first bending of the frame, a delivery truck could be modeled as a short frame carrying an external mass above it). For

this reason, gross size to the external observer does not always correlate with how moment is carried through a structure. For example, a railroad tank car relies heavily upon the tank itself to carry structural loads, while a highway truck has elastomeric mounts between the vehicle frame and cab that effectively decouple the system and prevent the cab from carrying significant strain energy during the first bending mode of the frame. The size parameter tabled here is for general comparisons; a more detailed look at each structure (and substructure) is recommended for a quantitative approximation of natural frequency using the parameters outlined here. DISCUSSION & CONCLUSIONS In summary, general effects of each parameter on the natural frequency of first beam bending are listed here in order of decreasing influence (recall that n = C end Csec Fmat Fsize ):

Fsize

for fixed aspect ratio, section & constraints, unlimited a 10x size decrease 10x fundamental frequency increase most structures within 200x bandwidth (i.e., 0.15-30 Hz)

Csec

practical factor up to 1.7 (different section modulii within a cross-section footprint) corners (rectangular sections) add 30% to bending frequency of same size cylinder intuitively second to size (e.g., cannot design a supertanker with f1st-bend 5 kHz)

C end

little difference between moving vehicle/vessel types (e.g., automobile vs. aircraft) factor of 2-3 difference between fixed and moving structures (railcar vs. building) worst-case factor of 6.4 (comparing fixed-fixed to fixed-free end constraints)

Fmat

little difference between metals (e.g., steel vs. aluminum) about 20% decrease from metals with fiberglass-polymer composites (same geometry) factor of 3-6 difference between ceramic and polymer extremes

Geometric size and aspect ratio clearly play critical roles in gross structural resonance. Although real structures behave in complex combinations of these parameters (and with variations in substructure modal participation), the role of material, section modulus and even boundary conditions are comparatively limited in driving fundamental frequencies for the first few modes of structural vibration. And despite their apparent differences, these more damaging gross flexible body modes5 for human-occupied structures typically occur at frequencies less than 30 Hz (this includes structures ranging in size from a sedan to a supertanker!) whether from the ground vehicle, aerospace, marine/offshore or civil engineering industries. Again, an article must shrink to the realm of small electronic components before significant damage mechanisms due to fundamental mode shapes jump a few orders of magnitude to the kilohertz bands (acoustics and noise concerns being notable exceptions). The implication for designers is that they will often not be able to solve problems excited by gross structural resonances through small geometry and/or material changes (i.e., it is difficult to shift the frequency of gross body mode shapes to a safe region of little input); they will often need to utilize damping to reduce the amplitude of any particular resonant response. Alternatively, designers can use increases in section modulus to reduce the resulting strains of an only slightly shifted natural frequency. The implication for structural test engineers is that they need only to sample digital data at rates high enough to capture mode shapes of interest, and do not need to collect at kilohertz rates (with the exception of impact or shock tests); setting low-pass anti-aliasing filters to about 30 Hz should be adequate to characterize most basic structural fatigue problems. The authors' practice is to base cut-off frequencies for an unfamiliar test setup on the estimated first natural frequency of the smallest substructural element of the primary structure under test.
5

The first few flexible body modes of vibration are often responsible for the larger bending and torsional stresses in a typical loading profile for a structure, due to the gross deflections and peak strains that accompany them. Higher-order shapes are often accompanied with smaller peak deflections and strains, and may sometimes be neglected during load collection for fatigue studies. This concept is explored further in the companion paper: Shust, W.C. and Smith, K.B., Bounding Natural Frequencies in Structures II: Local Geometry, Manufacturing and Preload Effects, SEM-IMAC XXII, 2004.

On a related note, maximum bending stress in a structure is also a function of the girth and squared length of an object but the relationship is inverted (i.e., max Lbeam 2 D ). So if the length of a structural design increases faster than the cross section (the apparent trend evidenced by new skyscrapers, SUVs and commercial jets), stresses may generally increase while natural frequency drops. Although this trend is limited (as Galileo originally observed) because the weight of a structure is proportional to its volume, or the third power of its size, while the strength of a structure is proportional only to its area, or the second power of its size. For this reason, a structural design that continues to grow in this manner will ultimately collapse under its own weight. The only known object that can grow indefinitely is a sphere. ACKNOWLEDGEMENTS, REFERENCES & BIBLIOGRAPHY The authors would like to express gratitude to Dr. Darrell F. Socie, for several water-cooler conversations and a particularly difficult homework assignment that inspired this exploration. This survey follows common issues regarding filtering choices for acquisition of mechanical test data [1]. First principles and basic mechanical properties for an array of common structural materials were reviewed [2-8]. Some representative structural aspect ratios and resonant vibration frequencies were collected from a mix of government standards [9-14], a sampling of vibration textbooks [15-22] and several technical papers from conference journals [23-37]. A brief internet survey [38-50] was also conducted to fill in some gaps.
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] Shust, W.C., Filtering Data to Better Understand Mechanical Systems, Experiences and Implications, Proceedings of 70th Annual Shock and Vibration Symposium, Albuquerque, NM, 1999. Socie, D.F., Failure Analysis of Mechanical Components, ME 347 Lecture Series, University of Illinois Board of Trustees, IL, 2001. Beer, F.P., and Johnston, E.R.Jr., Mechanics of Materials, 2nd Ed., McGraw-Hill, Inc., NY, 1992. Shigley, J.E., and Mischke, C., Mechanical Engineering Design, 5th Ed., McGraw-Hill, Inc., NY, 1989. Thomson, W.T., and Dahleh, M.D., Theory of Vibration with Applications, 5th Ed., Prentice Hall, Inc., NJ, 1993. Callister, W.D., Jr., Materials Science and Engineering, 2nd Ed., John Wiley & Sons, Inc., NY, 1991. "MatWeb", http://www.matweb.com, Online Material Property Database, Accessed April 2003. Metals Handbook - Desk Edition, American Society for Metals, OH, 1997. Earthquake-Induced Dynamic Response of Bridges and Bridge Measurements, NRC Transportation Research Board Record 579, DC, 1976. Environmental Engineering Considerations & Lab Tests: MIL-STD-810E, Naval Ship Systems Command, DC, 1974. Mech. Vibrations of Shipboard Equipment: MIL-STD-167, Naval Ship Systems Command, DC, 1974. Pressurized Rescue Module System, SOW-TM301103, Naval Sea Systems Command, DC, 1999. Transportation Characteristics of Truck, Rail and Water Freight, NRC Transportation Research Board Record 577, DC, 1976. Transportation Ride Quality, NRC Transportation Research Board Record 646, DC, 1977. Barra, R.J., Geo-Metric Vibration Analysis, RMS Publishing Co., MD, 1977. Iwan, W.D., Applied Mechanics in Earthquake Engineering, American Society of Mechanical Engineers, NY, 1974. Lomax, T.L., Structural Loads Analysis for Commercial Transport Aircraft: Theory and Practice, American Institute of Aeronautics and Astronautics, VA, 1996. McGoldrick, R.T., Ship Vibration, David Taylor Modal Basin Navy Report 1451, DC, 1960. National Research Council, Earthquake Engineering Research, National Academy Press, DC, 1982. Perry, M.A., Flight Test Instrumentation, Vols. 2 & 3, Pergamon Press, Great Britain, 1965. Todd, F.H., Ship Hull Vibration, Edward Arnold Publishers, Ltd., Ireland, 1961.

[22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47]

Vertes, G., Developments in Civil Engineering: Structural Dynamics, Elsevier Science Publishing Co., Hungary, 1985. Dynamic Waves in Civil Engineering, Society for Earthquake and Civil Engineering Dynamics, England, 1970. Earthquake Ground Motion & its Effect on Structures, ASME Applied Mechanics Division, AZ, 1982.
st 1 International Modal Analysis Conference, NY Union College, FL, 1982. th

11 International Modal Analysis Conference, Society for Experimental Mechanics & NY Union College, FL, 1993. Modal Testing and Analysis, ASME Technical Committee on Vibrations and Sound, MA, 1987. Modal Testing and Refinement, ASME Applied Mechanics Division, MA, 1983. Proceedings of Damping 89, Flight Dynamics Laboratory of the USAF Wright Aeronautical Laboratories, FL, 1989. Ship Vibration Symposium, Society of Naval Architects and Marine Engineers, VA, 1978. Sloshing and Fluid Structure Vibration, ASME Pressure Vessels and Piping Conference, HI, 1989. Use of Models and Scaling in Shock and Vibration, ASME Shock and Vibration Committee, PA, 1963. Vibration and Noise in Motor Vehicles, The Institution of Mechanical Engineers, England, 1972. Vibration in Civil Engineering, International Association for Earthquake Engineering, Britain, 1966. Vibrations in Rotating Machinery, The Institution of Mechanical Engineers, England, 1977. Vibration Control and Active Vibration Suppression, ASME Technical Committee on Vibrations and Sound, MA, 1987. Vibration Control in Microelectronics, Optics and Metrology, International Society for Optical Engineering, CA, 1992. "Boeing 747-400 Freighter Product Review," http://www.boeing.com/commercial/747family/flash.html, The Boeing Company, Accessed May 2003. "Determination of Natural Frequencies and Mode Shapes of Multi-Degree of Freedom Structures," http://ucist.cive.wustl.edu, University Consortium on Instructional Shake Tables, Accessed May 2003. "DMU (Passenger Railcar) Specifications Page," http://www.coloradorailcar.com /dmupages/dmuspec.html, Colorado Railcar Manufacturing, LLC, Accessed May 2003. "Earthquakes and Earthquake Engineering: FAQs," http://mceer.buffalo.edu/inforService /faqs/eqaffect.asp, Research Foundation of the State University of New York, Accessed May 2003. "Exploratorium: Life Along the Faultline," http://www.exploratorium.edu/faultline/engineering.html, Accessed May 2003. "GE Rail Services Equipment," http://www.ge.com/capital/rail/index.shtml, General Electric Railcar Leasing Division, Accessed May 2003. "Golden Gate Bridge Facts," http://www.thoma.com/thoma/ggbfacts.html, transcribed from CA Highway and Transportation District, Accessed May 2003. "Our Aircraft," http://cessna.com/aircraft/, Textron Cessna Aircraft Company, Accessed May 2003. "Our Final Project: The Golden Ratio," http://www.geom.umn.edu/~demo5337/s97b/, The University of Surrey, Accessed May 2003. "Resonance, Tacoma Narrows Bridge Failure and Undergraduate Physics Textbooks," http://www.ketchum.org/tacomacollapse.html, paper link from History of the Tacoma Narrows Bridge, Accessed May 2003. "Seabox ISO Containers & Container Specifications," http://www.seabox.com /CS_00_Container_Specs.html, Intermodal Container Survey, Accessed May 2003. "Simviation Model: C5 Galaxy," http://www.simviation.com/rinfolocc5.htm, World Flight Simulation & Aviation, Accessed May 2003. "Structural Engineering Beam Structures," http://www.aku.ac.ir/faculty1/aliniamm /Structural%20Slides/Beams/, Survey of Bridge Types, Accessed May 2003.

[48] [49] [50]

You might also like