You are on page 1of 18

STRUCTURAL SERVICEABILITY: FLOOR VIBRATIONS

By Brace Ellingwood, 1 M . ASCE a n d A n d r e w Tallin 2

ABSTRACT: Floor vibrations arising from normal human activity may affect the
serviceability of modern building structures, which are becoming lighter and
more flexible. Existing serviceability criteria for floors are reviewed in the light
of research dealing with human perception of structural motion. The dynamic
response of floors to realistic pedestrian movement excitation models is ana-
lyzed. Tentative serviceability criteria to minimize floor vibrations that are ob-
jectionable to building occupants are presented.

INTRODUCTION

Serviceability limit states in structural engineering occur when the


function of the structure is disrupted because of local minor damage,
deterioration of structural or nonstructural components, or excessive
structural movement (1,12,15,25,33). Excessive structural deflections are
a main source of unserviceability in buildings. Static deflections may af-
fect the appearance or efficiency of structural and nonstructural ele-
ments and mechanical equipment. Dynamic deflections may lead to oc-
cupant discomfort and impaired function of sensitive mechanical
equipment, especially if resonance occurs.
Structural vibrations arise from normal human activity and from the
operation of mechanical equipment within buildings, from external traffic,
or from wind storms and earthquakes. Structural vibrations under con-
ditions of normal use were not usually a problem when working stress
methods were used to design conventional floor and framing systems
with traditional construction. This can be inferred from the minimum
guidance provided in existing standards for controlling vibrations in or-
dinary buildings (e.g., Refs. 7, 30). However, structural systems are be-
coming lighter and more flexible and have lower damping than before.
There has been a reduction in the use of nonstructural members and
partitions which, in the past, provided additional stiffening and damp-
ing. Methods of structural analysis and design are growing more re-
fined, the systems are better integrated, and the use of high strength
construction materials and welded or bolted joints is now common (15,33).
Floors of light construction with longer spans afford cost advantages and
versatility in multi-purpose buildings. Limit states design, with its em-
phasis on structural behavior, also results, in some instances, in more
flexible structural systems.
These recent developments in design and construction practice may
lead to structural systems in which the motion is perceptible and objec-
'Research Struct. Engr., Center for Building Technology, National Bureau of
Standards,
2
Washington, D.C. 20234.
Research Asst., Dept. of Civ. Engrg., The Johns Hopkins Univ., Baltimore,
Md. 21218.
Note.Discussion open until July 1,1984. To extend the closing date one month,
a written request must be filed with the ASCE Manager of Technical and Profes-
sional Publications. The manuscript for this paper was submitted for review and
possible publication on February 24, 1983. This paper is part of the Journal of
Structural Engineering, Vol. 110, No. 2, February, 1984. ASCE, ISSN 0733-9445/
84/0002-0401/$01.00. Paper No. 18597.
401
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/cop
tionable unless it is controlled by proper design. Serviceability require-
ments in current standards and specifications, which are little more than
rules of thumb based on experience with traditional construction, often
are not sufficient for minimizing objectionable motion in flexible modern
structures.
Comprehensive and rational loading and resistance criteria have been
developed for the ultimate or safety-related limit states (13). However,
in many instances, serviceability requirements will determine the degree
to which limit states design will lead to feasible and economical struc-
tures. This paper reviews existing serviceability criteria for floors in the
light of research dealing with human perception of structural motion.
Following this review, tentative serviceability criteria to minimize floor
vibrations that are objectionable to building occupants are presented.

DEFINITION OF LIMIT STATE

Building occupants expect that the building will remain stationary un-
der ordinary conditions. When floors or the structural frame as a whole
vibrate unexpectedly, the occupants may become annoyed or alarmed.
People perceive structural movements that are far less than would cause
structural damage in a properly designed and detailed structure
(14,15,18,31). However, the average person does not appreciate this, and
is likely to take structural movement as an indication that the building
is unsafe. Recent publicized structural failures have reinforced this per-
ception. Such considerations suggest that vibration limits, at least for
buildings, should be related to levels of motion that building occupants
are willing to tolerate.
Numerous studies have been conducted in attempts to relate levels of
structural motion to human comfort or tolerance levels, as summarized
in Refs. 15 and 25. In general, these studies have involved subjecting
volunteers to a variety of low-amplitude excitations and having them
identify which vibrations are perceptible, tolerable, annoying, or uncom-
fortable. Relations have been established between peak or root-mean-
square (rms) displacement, velocity and acceleration and the fundamen-
tal natural frequency and damping in the structural system. Many of the
vibration tolerance limits thus developed have been tied to the 1931 Reiher-
Meister study (29), which involved steady-state excitation. Subsequent
modifications (e.g., Ref. 20) have attempted to account for the transient
nature of certain types of excitation.
The specific findings in these various studies are not always consistent
with one another because of the diverse purposes and methods in the
experiments. The psychological factors involved in structural motion
perception are difficult to duplicate in a laboratory experiment (25). The
willingness of experimental volunteers to tolerate higher motion levels
when they are forewarned and know no danger is involved suggests
that the tolerance and comfort levels reported from controlled experi-
ments may be upper bounds on the levels that may actually be accept-
able in an uncontrolled setting within a building. However, motion per-
ception levels generally are lower when people are expecting or are cued
for motion. Despite the uncertainty regarding the validity of the specific
motion tolerance levels reported, there are several general common trends
402
wnloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyrig
in the findings of these studies. Thus, some general conclusions can be
drawn that are useful in identifying structural motion serviceability limit
states.
One general finding is that larger amplitudes of transient motion that
dissipate within a few cycles are more easily tolerated than steady-state
motion that occurs over an extended period of time (11,15,23). The pe-
riod of decay of the vibration, which is determined by the damping, is
especially important for minimizing the effect of transient excitation.
Damping in floor systems is increased by the presence of floor cover-
ings, furniture, partitions, and people (1-3). If the motion essentially
damps out within approximately 5 to 10 cycles, the level of acceleration
considered tolerable is higher by approximately a factor of 10 than if the
excitation is continuous (20,21).
Structural acceleration appears to be the most appropriate response
quantity to use as a limiting motion parameter (12,18,23,31). The accel-
eration (and its derivative) is directly related to the whole-body forces
that are sensed kinesthetically. Annoyance threshold accelerations are

0.50
/
/
/
/
0.20 Transient /
(damping = 0.12) / A
/
0.10 / J
/
Transient /
(damping = 0.06) /'
0.05
a:

8 0.02 Transient /
(damping = 0.03) /

0.01

Continuous
0.005

0.001
5 10 20
FREQUENCY (Hz)

FIG. 1.Annoyance Threshold Vibrations for Residences, Offices and Schools


(After Refs. 1 and 22)

403
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/cop
shown in Fig. 1 (1). Whole-body vibrations in the range of 3-10 Hz are
particularly critical because large body organs within the rib cage and
abdomen resonate within this frequency range and it is within this gen-
eral range that the inherent vibration isolation capability of the body is
least effective (16). Vibratory movement at frequencies less than 1 Hz
may be associated more with motion sickness than with whole-body re-
sponse to dynamic forces. Motion in this frequency range is a special
consideration in minimizing the effect of wind excitation on very tall
buildings. The tolerance of a person to vibration also depends on his
activity at the time motion occurs, suggesting that limits should be re-
lated to the primary occupancy of the building.
Within the natural frequency range of approximately 1 to 8 Hz, the
acceleration corresponding to a threshold of annoyance is about 0.005 g
to 0.01 g (0.05 to 0.10 m/s 2 ) for continuous vibrations (1,2,15,29,35), in
which g is acceleration due to gravity (9.81 m/s 2 or 386 in./sec 2 ). People
engaged in quiet activities tend to find vibrations above 0.005 g objec-
tionable. For transient vibration in floors in which structural damping
is about 6% of critical, the annoyance threshold is about 0.05 g (0.50
m/s 2 ) (1,20). Note that amplitudes of vibration necessary to cause struc-
tural damage are generally well above these levels (31). With the excep-
tion of earthquake effects, few cases have been reported in which vi-
brations have been the likely cause of failure or severe structural damage
(15). Earthquake effects are covered under safety-related limit states and
are so infrequent that they are not a serviceability problem.
In contrast to the ultimate limit states where the definition of failure
is reasonably clear, there frequently is not a clear distinction between
what is serviceable and what is not. The transition into the structural
vibration serviceability limit state, where the level at which motion be-
comes objectionable depends on the individual and his activity when
motion occurs, is gradual and subjective. A recent study has considered
the limit state of excessive static deflection, and has treated the transition
to unserviceability using utility analysis (28). The serviceability criteria
developed using this approach were found to depend mainly on the
minimum or threshold deflection limit selected rather than on the range
of limits over which the utility function varied. Additional data are nec-
essary to define properly the utility functions used in this approach.
Meanwhile, it appears that neglecting the randomness in the parameter
that describes the limit state (acceleration in the present study) will not
affect the serviceability criteria developed to any great extent.

CURRENT DESIGN PRACTICE AND PROBLEM AREAS

Current codes and standards attempt to ensure that floor systems re-
main serviceable under conditions of ordinary use in basically two ways.
The first is by limiting the elastically computed static deflection 8 due to
nominal live load (4) to a simple arbitrary deflection limit which is some
percentage of the length of span, (. A common requirement is that 8
should be limited to /360; values ranging from /180 to /480 are cited
in different specifications in the United States (7,30). The second ap-
proach establishes a minimum depth for flexural members with respect
to their span. This minimum may depend on the end restraint (7) or on
404
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyr
the strength of the flexural member (30). For beams supporting large
open floor areas, it is recommended that member depth h should not
be less than /20 in order to minimize perceptible transient vibrations
(Ref. 30, Commentary). The static deflection limit of approximately /
360 appears to have been in use for over 100 yr, while the limitation on
i/h of 20 has been used as a vibration control measure since before 1914
(1,11,15). Construction practice obviously has changed considerably since
these simple limits were recommended.
These limits are an embodiment of professional experience with tra-
ditional structures and construction materials that, for the most part,
have performed satisfactorily in service. They have been used to control
minor nonstructural damage, objectionable structural movement and other
conditions of unserviceability. As long as structural bay sizes of 20 ft (6
m) or less were usual and design was based on working stress methods,
deflections or vibrations were seldom a problem. However, it is difficult
to point to any one state of structural behavior that such requirements
prevent. For example, limits on deflection as a percentage of span limit
the curvature and the maximum strain in bending of flexural members.
While this might prevent damage to ceilings, partitions and other at-
tached nonstructural components, such stiffness requirements may not
be helpful in minimizing human discomfort from structural motion. For
the latter, acceleration of the floor system, damping, and whether the
excitation is transient or steady-state are more important factors.
Current serviceability requirements have certain implications for the
dynamic response of flexural members. The flexural rigidity EI of a sim-
ple beam can be related to its fundamental natural frequency /x (in Hz)
by (9),

A =ie^ V
\ / PT W
in which p = (zvD + pwL)/g = the mass per unit length; wD and wL are
dead and live load per length; and p = the percentage of the nominal
live load that actually participates in the dynamic motion. For general
and clerical offices, the mean live load acting on the floor at any arbitrary
instant of time is about 11 psf (527 Pa) (8) while the nominal live load
is between 25 and'50 psf (1,197-2,394 Pa), depending on the loaded area
considered. Thus, p typically would be less than 0.25, since not all the
live load participates in the dynamic motion.
If the beam satisfies its static deflection limit of /360 under full live
load wL, this is tantamount to limiting its fundamental frequency to

U s 3.4 /-^2_f (2)

wD
For a typical wJwD = 2 for steel structures (13) and p = 0.25, Eq. 1
becomes fa s 22/V?, where i is measured in feet. Similarly, if the beam
is stressed to F in flexure under the effects of wD + wL, the fundamental
frequency is
405
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copy
.1 IL^J^L
2
{3a)
-/16F1 + ^

WD

If the beam is fully stressed, F = 0.60-0.66 times the yield stress Fy (30)
and thus 16 Fa ~ 10 F y . Moreover, suppose that (/h 800/Fy in accor-
dance with the commentary to Ref. 30, in which Fy and are in ksi units.
Then

1 + ^
f /_sL_S2.1 m
1
2/8,000 | W>LJ
V wD
which becomes /j s 24/V for the same p and wL/wD as before. Note
that structural frequencies obtained from these expressions are nearly
identical. If i = 26 ft (8 m), fx > 4.5 Hz while if i = 52 ft (16 m), >
3.2 Hz. As will be shown subsequently, these fa are close to the range
of excitation frequencies associated with several common human activ-
ities (2), Thus, the current deflection limit or span-depth limits might
not be adequate to control objectionable vibrations for spans much in
excess of 30 ft (9 m).
Floor vibration usually is not a problem in residential construction.
The fundamental frequencies of vibration are in the range of 15 to 30 Hz
(5,11,26) where tolerable acceleration levels are higher. Moreover, the
damping is quite high, on the order of 12%-15% of critical (26), and mass
of the floor is small enough that the people involved provide additional
damping. There is some evidence (11) that in light residential construc-
tion where the spans are typically 10-15 ft (3 to 4.6 m), a static deflection
limit of /360 under the 40 psf (1.9 kPa) nominal live load (4) results in
an acceptable floor.
Floor vibration problems are more likely to occur in offices and public
assembly occupancies where spans of 30 ft (9 m) or more are becoming
common and where there are large open areas without floor coverings
or partitions to augment damping (1,2,3,15,20,21). Beam-slab floor sys-
tems within steel frames that have experienced vibration problems gen-
erally have had floor beams with /h greater than 20 or concrete slabs
less than 4 in. (101 mm) thick (21). Such floor systems have fundamental
structural frequencies in the range 4-10 Hz. Floors and walkways in as-
sembly areas where the common activities are such that crowds move
to a beat are particularly susceptible to excessive dynamic motion. Per-
sons at spectator events may create forces with frequencies in the range
of 2-4 Hz (2). Many structural systems for assembly occupancies have
relatively long spans and natural frequencies that are also in this range.
Substantial structural accelerations (greater than 0.3 g) have been mea-
sured during sports events and rock concerts (2).
Considerable research on the performance of light joist-slab floors has
been performed (e.g., Refs. 20, 21). These studies showed that theoret-
ical predictions of dynamic behavior of floor systems correlate well with
406
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyr
observed behavior of actual floors loaded dynamically. Shear friction be-
tween joists and slab was found to be sufficient to cause composite ac-
tion under small amplitude excitation. The rate of decay of structural
response (damping) was a particularly important indicator of human
sensitivity to motion. This research has led to some preliminary design
recommendations and portions have been adopted for at least one de-
sign specification (22). However, the conclusions of this research are all
derived from using a heel drop impact test, in which a person supports
his weight on the balls of his feet with the heels raised approximately
2.5 in. (64 mm) and then relaxes, allowing his heels to impact the floor.
The heel impact test may be meaningful for evaluating the response of
floor systms to activities that cause impact forces (e.g., jumping in place
or dropping heavy objects). It appears, in many cases, to provide a use-
ful tool for screening some potentially troublesome floors (1,20). How-
ever, the heel impact test produces an isolated transient vibration which
is not a good simulation of vibrations due to walking, which are a com-
mon annoyance. The heel impact test indicates a stronger dependence
of the acceptability of the floor on damping than has been found in other
research (35). The vibrations produced by walking tend to merge to-
gether and produce nearly continuous motion of the floor.
Other studies of floor vibrations have considered the response of a
floor to a walker treading in place (6) or have assumed purely sinusoidal
excitation (12) to determine peak dynamic response. These assumptions
may also be unrealistic. Treading in place causes an essentially steady-
state response and does not account for the fact that the walker excites
the floor with a different intensity depending on his position as he moves
across the floor. Such force representations may not be sufficient for
establishing general performance requirements for floor systems. Forc-
ing functions that model other activities should also be considered, as
described subsequently.

FORCING FUNCTIONS

Normal floors usually are not exposed to steady-state excitation from


machinery. If they are, the most effective way to deal with the problem
is through use of vibration isolation devices. The main sources of floor
vibration are activities of human occupants. A brief summary of forces
arising from such activities follows. Effects of walking are particularly
important because of the frequency with which they occur.
Forces Due to Walking.Several studies of forces exerted on a floor
by normal walking activity have been conducted for such widely varied
purposes as evaluating resistance to abrasion in floor surfaces (17) and
designing prosthetic devices (19). A typical nondimensionalized force-
time relation for a single footfall is shown in Fig. 2(a). The first and sec-
ond peaks correspond to heel strike and toe liftoff contact, respectively.
A Fourier analysis (19) of the force pulses due to normal walking shows
that the first and third harmonics of the pulse are most significant. The
average relative magnitudes of the second and third harmonics are ap-
proximately 20% and 45%, respectively, of the first harmonic, while the
harmonics of four and above contribute less than 10%. As the individ-
ual's gait increases from a walk to a run, the peaks in Fig. 2(a) merge

407
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyr
,.20
BOO
V 1-23 J^-V - 1 . 1 5
700
^-OM \ 1.00
000
0.75
g1 ~
SOO
400
\
300

200
0.25
100
0 1() . <, '
lb)
( H

FIG. 2.Forees Due to Walking on Floor: (a) Foree from Slngte Footfall; (h) Forces
on Span from One Person Walking

together and the peak forces approach roughly 2.75 times his body weight
(24,34). Statistical analysis of force data in Ref. 17 suggests that the peak
forces shown in Fig. 2(a) have a probability of approximately 0.005 of
being exceeded. The normal rate of walking is about 112 steps/min, cor-
responding to a forcing frequency of about 2 Hz. The excitation of a floor
system by one person walking is shown spatially and temporally in Fig.
2(b). The force pulses due to successive footsteps will normally overlap
by roughly 0.1 sec.
Dynamic forces from groups of people walking about the floor at ran-
dom would seldom cause serviceability problems. The static deflection
of the floor naturally would increase. However, the excitation would
lack coherence unless the group was walking in step, and thus the dy-
namic component of motion would be small. Others (34) have also con-
cluded that the single pedestrian provides an appropriate excitation model
for developing serviceability criteria.
Forees Due to Dancing, Jumping, and Other Rhythmic Activities in
Place.These forces arise from the response of a group of people to
some external stimulus such as music or a spectator event. The forces
are distinguished from forces due to walking by the assumption that the
spatial variation in the force is negligible in comparison to the temporal

1 1" "I
EVENT Ha. t|HC|
r\-F[ioi Stumbling 2.6 0.02
000 Filling
ACTIVITY a l|h;| backward 2.0 0.03
Dancing lump tram
1.0 2 -3
2 I t |0.0 m| 7.6 0.04
Running or
a . 1 / ,-W-r...lMI lumping In
Jumping in plico 2.7 2 - 4
PI9C9 3.9 0.03
Cheering 0.7 2 -4 |llt| .
Walking In place 0.8 2 J - I I \ \
'Normalized to static
lorco F,

1 i

f\ i \ i i
1 OF, 1
aF,
200
r N K t s l drop Impact
\ t l l t |20|
/.' Tiptoe-to-heel \

V \\JI [J \ < -twr II


/
rmpsct.ons toot \
at i Mm |24| " \
I
Fj = Static torca \ '
0
\ 1 \ '
0.02 0.03
- t[aec]

FIG. 3.-Forces Caused by Rhythmic FIG. 4.Forces Due to impact Events


Activities In Plaee (Refs. 2,12, 24) (Rets. 20, 24)

408
ownloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyri
variation. The general characteristics of these forces are quite similar. A
sample force function is shown in Fig. 3 (2,12,24). The factor a by which
the static force is multiplied to obtain the dynamic force component var-
ies over a considerable range, as shown in the figure.
Forces Due to Impact Events.The distinguishing feature of these forces
is that their duration is usually short in comparison with the funda-
mental period of vibration of the floor. Such forces might arise from
people jumping off furniture or ladders or dropping heavy objects. Force
versus time relations inferred from results of tests that simulated several
impact events (24) are shown in Fig. 4. Included for comparison is a
force record measured from the heel drop impact test that has been used
extensively in previous studies of floor vibration (1,20,22).

RESPONSE OF FLOOR TO WALKING

The predicted responses of simple floor systems to normal walking,


to sinusoidal excitation and to a heel drop are compared in this section.
Assume that the floor behaves as a one-way flexural member and that
its fundamental mode dominates response. The force /(), 0 < t < te,
for a single footfall is given in Fig. 2(a). The fundamental mode shape
of the floor is <t>i(y) and its corresponding natural frequency is ^ . The
modal response to a single foot fall at point y on the floor span is

*iW = Hy) fWHt - T)rfT (4)

in which h(t - T) = sin cox D(t - T) (5)


0) 1D Mti
in which i = t when 0 < t < te, i = te when t > te; a)iD = cof(l - (3?);
mx and fa are the mass and damping in the fundamental mode.
When a man walks across the floor taking l/ts steps per unit time with
a stride length of Ay, the response of the floor during interval (n - 1) ts
< t s n ts is

xi(t) = Hyi) fWHt ~ t)dT + Hyi + Ay) / W M T - (f - t,)]dj + ...


Jo Jo
/f-(M-l)t s

+ <Myi + ( - l)Ay] /(T)/I{T - [t - (n - l)t,]}dT (6)


Jo
in which yx = the coordinate of the first step. Letting

X(f) = J /(T)fe(T - t)dt (7)


Jo
with t defined, one has for the interval (n - l)fs < t < nts
n

*iW = E <Myi + 0' - DAy]x[ - (i - i)t,].' (8)

409
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyr
Eq. 8 can be modified when the effect of higher modes is to be con-
sidered or when the floor is assumed to behave as a plate rather than
as a simple beam. The deflection at some point y is then

*(*'# = I >=1 J'-l


<j>,(i/i + ( / - l ) A i / ) X 1 . ( t - ( / - l ) t s ) (9)

in which <> | , = the shape of the z'th mode for the floor system considered;
m = the number of modes; and X, = the Duhamel integral for the ith
mode (9) (Eq. 7). A computer program has been written to perform the
calculation in Eq. 8 for a floor assumed to be in one-way bending with
the familiar mode shapes <(>, = sin (itry/() in which = the span and i
= 1, 2, 3. The program numerically integrates Eq. 7 for an arbitrary foot-
fall force versus time record.
Consider for illustrative purposes a one-way floor spanning 26 ft (8
m) with modal frequencies of 5, 20, and 45 Hz for the first 3 modes.
The structural damping fa is 3% of critical. The calculated static deflec-
tion of this floor system under a uniformly distributed live load of 50
psf (2.4 kPa) is 0.47 in. (12 mm), or /665, so the floor would satisfy the
deflection limit in existing codes. The mass of the walker is assumed to
be small in comparison with the mass of the floor.
The assumed forcing function for a single footfall shown in Fig. 2(a)
has been scaled to describe the force exerted on the floor by a single
footfall of a 157 lb (700 N) person walking at a moderate pace (approx-
imately 2 paces/s) (17). An examination of the modal responses to pe-
destrian movement indicated that the first mode dominates the displace-
ment response and the third mode is more significant than the second.
This would be expected from the shape of the forcing function in Fig.
2(a). Although the accelerations in the higher modes can be quite high,
they damp out rapidly and lack coherence. Thus, the root-mean-square

0.0015

.1.5 2 2.5 3 3.5


TIME (sec)

FIG. 5.Response of 8-m Floor to Pedestrian Movement (/, = S Hz, ft = 0.83)


410
ownloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyr
1
Li, St
? i . ArtN ff1/
m
w
m
o
Ul
0015

|V V r1
K
(V
0
I 2.5 3 3.5
0.5 1 1.5 2 2.5 3 3.S
TIME (sec)

FIQ. 6.Response of 8-m Floor to Pedestrian Movement: (a) /, = 10 Hz, fr = 0.03;


(b)/! = 10 Hz, (3i = 0.12

(rms) accelerations in these higher modes are substantially smaller than


the rms acceleration in the fundamental mode. Since human perception
of motion is known to depend on the length of time over which signif-
icant acceleration is experienced, one may conclude that the fundamen-
tal mode response is the major contributor to human discomfort.
Fig. 5 shows the predicted response of the floor system to pedestrian
movement [Fig. 2(b)]. The static as well as dynamic components of de-
flection are clearly evident. The dynamic response with the walker on
the center half of the span approaches a steady-state conditions with a
maximum acceleration of 0.026 g. The acceleration of this particular floor
is governed by the third harmonic of the footfall, which has a frequency
of about 4 Hz. In fact, a sinusoidal excitation with frequency of about 4

0 0.5 1 1.5 2 2.5 3 3.5 4.5 5


TIME (sec)

FIQ. 7.Response of 16-m Floor to Pedestrian Movement (/j = 10 Hz, PJ = 0.03)

411
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copy
Hz (about twice the normal rate of walking) and dynamic force factor a
of about 0.2 (rather than a = 0.5 to 0.6, as in Fig. 3) would cause a
similar response as that in Fig. 5. The essentially steady-state response
suggests that damping in the floor may not be as significant a parameter
in assessing serviceability (at least for this common activity) as some
studies (20,21) indicate. However, damping obviously would be impor-
tant in controlling response to less frequent impact events. Using the
walking sinusoidal excitation suggested in Fig. 3, the acceleration of the
floor would be 0.0079 g.
Figs. 6(a)-6(b) show the response of a similar 26-ft floor system for
which the fundamental natural frequency /j has been increased to 10 Hz-
by increasing its stiffness. The dynamic response is smoothed by in-
creasing pi from 0.03 to 0.12 [Figs. 6(a) and 6(b), respectively], but the
displacement amplitudes are nearly the same for both damping ratios.
The computed peak accelerations are in the range 0.004 g-0.006 g, which
borders on perceptibility. In Fig. 7, the span of the floor has been in-
creased to 52 ft (16 m), while fx is maintained at 10 Hz and Pi is 0.03.
The accelerations in this stiff, massive floor system are approximately
0.002 g, well below the thresholdof perception.
The responses of the floor systems analyzed were found to be sensi-
tive to the shape of the footfall force pulse. Fig. 8 shows the response
of the floor spanning 26 ft (8 m) with fx = 5 Hz and Pi = 0.03 to the
slightly modified footfall pulse shown in the figure insert. The steady-
state acceleration is approximately 0.012 g, about one-half the accelera-
tion indicated for the same floor system in Fig. 5. The span has been
increased to 52 ft (16 m) and /i = 5 Hz in obtaining the results in Fig.

0 .05 0.1 0.1 S 0.J 0.25 0.3 0.35 0.4 OAS 0.5
J | TIME (SBC) |

0.5 1 1.5 2 2.5 3.5

TIME (sec)

FIG. 8.Response of 8-m Floor to Modified Footfall Pulse (/i = i Hz, fa = 0.03)
412
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copy
FIG. 9.Response of 16-m Floor to Modified Footfall Pulse (fx = 5 Hz)

9. The accelerations are about 0.011 g for damping in the range 0.03-
0.12. The sensitivity of the response to the shape of the footfall force
pulse indicates that accelerations predicted using simplified represen-
tations of the footfall force pulse may not be reliable. Additional data
on these forces obviously would be desirable.
The peak accelerations at midspan caused by the sinusoidal excitation
in Fig. 3 corresponding to walking in place are approximately one-third
the accelerations that would be caused by normal pedestrian movement,
while the accelerations from a heel drop (Fig. 4) at midspan range from
8 to 100 times larger. The range in these accelerations, when plotted on
a vibration tolerance chart such as Fig. 1, is sufficient to cause an ac-
ceptable floor system to be rated as unacceptable. This analysis confirms
the observation, communicated to the writers by several practicing en-
gineers, that no particular problems have been encountered with several
floor systems that would be classified as unacceptable according to a
heel drop test. Thus, the use of realistic force functions is important in
assessing the sensitivity of floors to disturbing dynamic motion.

DEVELOPMENT OF SERVICEABILITY CRITERIA

Simple criteria can be developed to control objectionable vibrations that


are caused by normal activities of building occupants, and rhythmic ac-
tion of jumping, dancing, and cheering at sports events and special events
such as rock concerts (2). Motion criteria related to the operation of spe-
cial equipment are difficult to establish because they depend on speci-
fications and operating characteristics of the equipment. The following
developments do not consider such criteria.
413
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyr
Activities such as jumping off stepladders or accidentally dropping
heavy objects cause the largest initial acceleration response in a floor
system. However, these events may occur only rarely and the vibrations
damp out quickly. Vibrations caused by walking are an everyday oc-
currence and may approach a steady-state condition under which the
vibration tolerance levels are known to be much less. Vibrations caused
by walking and dancing are known to be more disturbing than transient
vibrations due to heel drop or other impact events, particularly in res-
idences (5,26). Thus, the floor vibration limit states criteria should ad-
dress two possibilities: low-level vibrations that occur frequently and large
transient vibrations that occur infrequently. Floors on which people move
about more or less at random, for which resonant amplification would
be unlikely, should be distinguished from floors in which people move
to a beat. While resonance would seldom cause problems in housing and
office occupancies, it may be a problem for sports arenas, dance halls,
and similar public assembly structures used for special events.
Serviceability criteria for floor vibration are derived from the perfor-
mance requirement that building occupants should not be disturbed un-
duly by floor movement under normal conditions and activities. (This
performance requirement is as far as most current standards go.) Sat-
isfaction of this performance requirement entails limiting the accelera-
tion of the floor to some tolerance threshold. This threshold should de-
pend on whether the motion is transient (damps out within a few cycles)
or steady state. Tolerable levels of acceleration should also be related to
the occupancy of the building (18). Hospitals would be most critical, fol-
lowed by residences, then offices, and finally industrial or other occu-
pancies where the primary activities within are of a physical nature and
provide distraction from motion.
The simple tabulation in Table 1 of acceleration limits for steady-state
motion for several of the general occupancy classifications in the A58
Standard (4) is based on the review of the literature on human tolerance
of motion summarized earlier. Table 1 represents a synthesis of the re-
sults of many studies. Definitions of limit states that are more complex
than simple acceleration limits would be very difficult to deal with in
routine structural design. Acceleration limits for transient excitation de-
pend on the degree of damping present, and are an order of magnitude
higher than those for steady-state excitation. While the acceleration lim-
its in Table 1 do not depend on frequency (cf. Fig. 1), the likely range
of frequencies for each occupancy has been considered in setting the
limits.
Serviceability of a specific floor system could be checked by calculating
its dynamic response to a suitable prescribed excitation. This checking
procedure would be difficult to use in routine design, which usually is
based on static calculations. Moreover, it would require the specification
of several excitation functions. Attempts at specifying force-time rela-
tions for designing against other dynamic effects have not been good
up to now (4). Thus, it is reasonable to expect that serviceability checks
will continue to rely on relatively simple rules that utilize static analysis
techniques.
If the excitation- of the floor is essentially sinusoidal due to rhythmic
activity with frequency /, the maximum acceleration at midspan for a
414
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copy
TABLE 1,Suggested Acceleration Limits by Occupancy
ACCELERATION LIMIT (g)
Frequent
steady- Infrequent Transient
state RMS Peak
Occupancy /activity (4) acceleration Damping acceleration
(D (2) (3) (4)
Hospitals 0.002 3 0.005
Laboratories 6 0.01
12 0.02
Hotels and multi-family apartments 0.005 3 0.02
Offices 6 0.05
School rooms 12 0.10
Libraries
Dining rooms and restaurants 0.01 3 0.05
Assembly areas, theaters 6 0.10
One- and two-family dwellings 12 0.20
Stores and shopping centers 0.02 3 0.05
Manufacturing, warehouses 6 0.10
Walkways, stairs and exitways 12 0.20
Dance halls and ballrooms
Recreational areas
Gymnasiums 0.05 N.A.
Stadiums and arena bleachers
Special assembly structures

lightly damped system is given (9) by


aF,
2 (10)
:~W) hr

in which aFs = the sinusoidal dynamic force; and/i = the fundamental


natural frequency of the floor. The term aFs/k is simply the static de-
flection due to force aFs. Using Eq. 10 and the steady-state acceleration
limits in Table 1, criteria to minimize objectionable floor vibrations due
to rhythmic activities take the form of a simple check on maximum static
deflection. As an example, if /i = 5 Hz, / = 4 Hz and a = 0.2, corre-
sponding to the third harmonic of normal walking, and xmax = 0.049 m /
s2 (0.005 g), the maximum permissible deflection due to aFs is 0.03 mm.
If /i increases to 20 Hz and all other parameters remain the same, the
maximum permissible deflection due to aFs increases to 0.07 mm.
This deflection is small because the term a.Fs is much less than the full
unfactored live load intensity wL (4) that is used in conventional deflec-
tion checks. Since the system remains elastic under serviceability con-
ditions, the force and deflection can be scaled upward toward the range
more common in conventional analysis. Considering several floor sys-
tems, the serviceability criteria for the second occupancy group in Table
1 might appear something like:
415
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copy
To ensure that building occupants shall not be unduly disturbed
by floor movement under normal conditions, the static deflection
of the floor under a 2 kN (450 lb) force applied at midspan shall
not exceed 1 mm (0.04 in.).

Similar requirements could be derived for the other groups in Table 1.


The peak acceleration from events that cause an impulsive force is
approximately

^it1" (U)
in which M = 0.5 p and impulse Im = 68 Ns from Fig. 4. Serviceability
for these events can be checked by setting a0 equal to the acceleration
limit given by occupancy in the second column of Table 1. Limiting the
natural frequency defined in Eq. 11 leads to a limit on beam rigidity of

EI a0
(12)
9 < ~2>n2In
3 6

It was assumed in the developments above that the floor behaves as


a simply supported flexural member. The dynamic stiffness of the floor
may not be increased by making it continuous over internal supports.
However, its static flexural capacity is increased and the section required
for a prescribed design load may be smaller. The decrease in the natural
frequency of the floor may increase the possibility of resonance with
excitation caused by human activities. This situation has led at least one
writer (21) to suggest that floor members should be designed as if they
were simply supported to ensure that their natural frequencies are well
removed from a resonance condition.
Floor systems with unsatisfactory dynamic response characteristics can
be modified in design through adjustments to their stiffness, damping
or mass. Increasing the stiffness and thus the frequency of the floor sys-
tem may not always be an effective way of reducing objectionable ac-
celerations. Increasing the stiffness may increase the separation between
the excitation and structural frequencies and reduce the possibility of
resonance. Moreover, at higher frequencies, the acceleration tolerance
limits tend to increasesee Fig. 1. If the response is essentially har-
monic, doubling the stiffness reduces the amplitude of vibration by a
factor of two. However, the square of the frequency is increased by a
factor of two, so the peak acceleration (Eq. 10) is unchanged. Increases
in mass are effective in reducing dynamic response but are difficult to
achieve because of economic restraints and space utilization require-
ments. A simple topping of 1 in. (25 mm) or less thickness on a slab
may be sufficient in some cases. Problems associated with transient ex-
citation may be dealt with by increasing the damping in the floor sys-
tem. As seen, however, many of the problem excitations in residences,
offices and assembly areas induce an essentially steady-state response.

CONCLUSIONS

The complex phenomenon of human perception of structural motion

Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copy
promises to become an increasingly important consideration in struc-
tural design. Limitations in existing specifications on the m a x i m u m de-
flections for flexural members or on their span-to-depth ratios are not
adequate for ensuring that objectionable structural motion caused b y
common occupant activities will not occur. Floor vibration problems fre-
quently can be minimized by requiring only a simple static deflection
check, in which the deflection limit d e p e n d s o n occupancy but is in-
dependent of span. In contrast, deflection criteria in existing specifica-
tions do not d e p e n d on occupancy; however, these latter limits m a y still
be necessary to minimize the possibility of d a m a g e to attached nonstruc-
tural elements.

ACKNOWLEDGMENT

The assistance of T. A. Reinhold during the early phase of the study


and comments offered by D. E. Allen a n d T. V. Galambos during the
preparation of the manuscript are gratefully acknowledged.

APPENDIX.REFERENCES

1. Allen, D. E., and Rainer, J. H., "Vibration Criteria for Long-Span Floors,"
Canadian Journal of Civil Engineering, Vol. 3, No. 2, June, 1976, pp. 165-173.
2. Allen, E. E., "Floor Vibration," presented at Venezuelan Concrete Confer-
ence, Caracas, Oct. 26-30, 1981.
3. Allen, D. E., "Vibrational Behavior of Long-Span Floor Slabs," Canadian Jour-
nal of Civil Engineering, Vol. 1, No. 1, Sept., 1974.
4. American National Standard A58, Minimum Design Loads for Buildings and
Other Structures (ANSI A58.1-1982), American National Standards Institute,
New York, N.Y., 1982.
5. Atherton, G., Polensek, A., and Corder, S., "Human Response to Walking
and Impact Vibration of Wood Floors," Forest Products Journal, Vol. 26, No.
10, Oct., 1976, pp. 40-47.
6. Becker, R., "Simplified Investigation of Floors under Foot Traffic," Journal of
the Structural Division, ASCE, Vol. 106, No. ST11, Nov., 1980, pp. 2221-2234.
7. "Building Code Requirements for Reinforced Concrete (ACI Standard 318-
77)," American Concrete Isntitute, Detroit, Mich., 1977.
8. Chalk, P., and Corotis, R. B., "A Probability Model for Design Live Loads,"
Journal of the Structural Division, ASCE, Vol. 107, No. ST5, May, 1981, pp.
857-872.
9. Clough, R., and Penzien, J., Dynamics of Structures, McGraw Hill Book Co.,
Inc., New York, N.Y., 1975.
10. Crist, R., and Shaver, J. R., "Deflection Performance Criteria for Floors,"
National Bureau of Standards Technical Note 900, Washington, D.C., Apr.,
1976.
11. "Deflection Characteristics of Residential Wood-Joist Floor Systems," Hous-
ing Research Paper No. 30, Housing and Home Finance Agency, Washing-
ton, D.C., Apr., 1954.
12. "Deformation Requirements for Buildings," National Research Council Tech-
nical Translation TT-1969, Ottawa, Ontario, Canada, 1980.
13. Ellingwood, B., Galambos, T. V., McGregor, J. G., and Cornell, C. A., "De-
velopment of a Probability-Based Load Criterion for American National Stan-
dard A58," National Bureau of Standards Special Publication 577, June, 1980.
14. Galambos, T. V., "Vibration of Steel Joist-Concrete Slab Floors," Steel Joist
Institute Technical Digest No. 5, Arlington, Va., 1974.
15. Galambos, T. V., Gould, P. L., Ravindra, M. K., Suryoutomo, H., and Crist,
R. A., "Structural DeflectionsA Literature and State-of-the-Art Survey,"
417
Downloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/cop
National Bureau of Standards Building Science Series 47, Washington, D.C.,
Oct., 1973.
16. Grether, W. F., "Vibration and Human Performance," Human Factors, Vol.
13, No. 3, 1971, pp. 203-216.
17. Harper, F. C , Warlow, W. J., and Clarke, B. L., "The Forces Applied to the
Floor by the Foot in Walking," Building Research Station, National Building
Studies Research Paper No. 32, HMSO, London, U.K., 1961.
18. Irwin, A. W., "Human Response to Dynamic Motion of Structures," Struc-
tural Engineer, Vol. 56, No. 9, Sept., 1978, pp. 237-244.
19. Jacobs, N. A., Skorecki, J., and Charnley, J., "Analysis of the Vertical Com-
ponent of Force in Normal and Pathological Gait," Journal of Biomechanics,
Vol. 5, 1972, pp. 11-34.
20. Lenzen, K. H., and Murray, T. M., "Vibration of Steel Beam-Concrete Slab
Floor Systems," Engineering Mechanics Report No. 29, University of Kansas
Center for Research and Engineering Science, Lawrence, Kan., 1968.
21. Lenzen, J. H., "Vibration of Floor Systems of Tall Buildings," Technical Com-
mittee No. 17 State-of-the-Art Report No. 4, in Planning and Design of Tall
Buildings, ASCE/IABSE, 1972, Vol. 2, p . 667.
22. "Limit States Design Steel Manual," Canadian Institute of Steel Construc-
tion, Willowdale, Ontario, Canada, Jan., 1977.
23. Nelson, F. C , "Subjective Rating of Building Floor Vibration," Journal of Sound
and Vibration, Oct., 1974, pp. 34-37.
24. Nilsson, L., "Impact Loads Produced by Human Motion," Swedish Council
for Building Research Document D13:1976, Stockholm, Sweden.
25. "Planning and Design of Tall Buildings," a monograph in 5 volumes, pre-
pared by the Council on Tall Buildings, ASCE, 1980, Vols. PC and SB.
26. Polensek, A., "Human Response to Vibration of Wood-Joist Floor Systems,"
Wood Science, Vol. 3, No. 2, Oct., 1970, pp. 111-119.
27. Rainer, J. H., and Allen, D. E., "Vibration Measurements on a Steel Joist
FloorMontreal," Division of Building Research Technical Note 573, Na-
tional Research Council of Canada, Mar., 1973.
28. Reid, S. G., and Turkstra, C. J., "Codified Design for Serviceability," Report
ST81-6, Department of Civil Engineering, McGill University, Montreal, Can-
ada, Dec, 1981.
29. Reiher, H., and Meister, F., "The Sensitivity of Humans to Vibrations," For-
schung auf dem Gebiete des Ingenieurwesens, Vol. 2, No. 11, Nov., 1931.
30. "Specification for the Design, Fabrication, and Erection of Steel Buildings,"
American Institute of Steel Construction, Chicago, 111., 1978.
31. Steffens, R. J., "Some Aspects of Structural Vibration," Current Paper En-
gineering Series 37, Building Research Station, Garston, U.K., 1966.
32. Vannoy, D., Barton, F., and Heins, C , "Design of Building Floor System
Subjected to Steady-State Dynamic Loads," Methods of Structural Analysis,
American Society of Civil Engineers, New York, N.Y., 1976, pp. 720-738.
33. Warwaruk, J., "Deflection RequirementsHistory and Background Related
to Vibrations," Vibrations of Concrete Structures, ACI Publication SP-60, Amer-
ican Concrete Institute, Detroit, Mich., 1979, pp. 13-42.
34. Wheeler, J. E., "Prediction and Control of Pedestrian-Induced Vibration in
Foot Bridges," Journal of the Structural Division, ASCE, Vol. 108, No. ST9,
Sept., 1982, pp. 2045-2065.
35. Wiss, J. F., and Paramelee, R. A., "Human Perception of Transient Vibra-
tion," Journal of the Structural Divison, ASCE, Vol. 100, No. ST4, Apr., 1974,
pp. 773-787.

418
ownloaded 08 Jun 2009 to 128.46.174.124. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyrigh

You might also like