You are on page 1of 5

compvtas a smc1n?es Vol. il. pp.

349-353
Pcrgamon Press Ltd., 1980. Printed in Great Britain

FLOATING ROOF ANALYSIS AND DESIGN


USING MIN~COMPUTERS~
HOWARD
I. EPSTEINS
Department of Civil Engineering, University of Connecticut, Storrs, CT 06268,U.S.A.
(Received 26 October 1978;received for publication 3 April 1979)
A~~-~oati~
roofs are used in oil storage tanks to reduce evaporation and hand&g losses, decrease
corrosion, and reduce the fire hazard. A pontoon roof is used in this paper. It consists of a circular centrai plating
attached at the edge to a compartmented; buoyant ring. The primary design Ioadings for this roof are due to
accumulated rainwater or a punctured deck. The design loads typically cause the deck to deflect several hundred
times its thickness and consequently the deck is usually designed and analyzed as a membrane.
The complicated nature of the loading on the membrane together with the unusual boundary conditions require
numerical integration of the coupled, non-linear differential equations which govern the behavior. A skilled analyst
is needed to vary the starting parameters necessary for the integration. This paper describes the procedure that
should be followed in the analysis and design of the roofs, and shows why an interactive computer is so impo~nt
in the understanding of the forces in this structure and in the design necessary to transmit them.

NOMENCLATURE

pontoon cross section area


lengths defined in Fig. 1
central water depth
Youngs modulus
edge oil head
edge water head
oil head given by HO-H,
horizontal pontoon stiffness
membrane radial force/unit iength
nondimension~ radial
force (eqn 4)
pressure on membrane
nondimensional pressure (eqn 4)
volume of rainwater
edge oil pressure
radius to centroid of pontoon
radius to element of membrane
nondimension membrane radius (r/d)
thickness of memb~ne
radial dispiacement of membrane element
vertical displacement of membrane element
nondimensional vertical displacement (w/t)
vertical displacement of membrane at r = 0
edge rotation (see Fig. 1)
horizontal edge displacement (see Fig. 1)
Poissons ratio
oil density
rainwater density
radial membrane central stress
radial membrane edge stress
pontoon rotation
INTROLWCTION
been used in structural

design
offices to assist with problem setups, to check the input
to programs run on large computers, and in some cases,
to solve entire problems. Minicomputers and time-shared
systems are both gaining in popularity because of their
relatively low cost. 30th have certain advantages, and
users must weigh their relative merits before investing in
one of the systems. Some tirms are pu~hasing minicomputers to take part of the burden off of their timeMinicomputer

have

tpresented at the American Society of Civil Engineers Convention and Exposition, Chicago, Illinois, K-20 October 1978.
SAssociate Professor.

sharing facility. The role of minicomputers in structural


analysis and design is changing rapidly; they are being
used for problems of ever-increasing complexityIf, 23.
As a result, many firms are ou~owing ~uipment purchased only a few years ago[3].
The particular application of an interactive minicomputer to the analysis and design of pontoon floating roofs
is presented in this paper. Floating roofs are installed in
oil storage tanks primarily to reduce evaporation and
h~dling losses, to decrease corrosion, and to reduce the
fire hazard. Boating roofs may be of the pan or double
deck type as well as the pontoon roofs considered
herein[4-6]. The pontoon floating roof consists of a
circular central plating attached at the edge to a compartmented buoyant ring (pontoon).
Over the years, the pontoon roof has experienced
structural problems as evidenced by several roof buckling and sinking faihrres. These failures are usually due
to either the accumulation of excessive rainwater on the
deck or the product leaking through punctures of the
deck or pontoon compartments. The design of these
roofs is typically governed by the forces produced by
one of these loadings.
In recent pubIications, the author and J. Buzek have
investigated the stresses and degections in pontoon
floating roofs as caused by the accumulation of
rainwater [7] or due to the product leaking through punctures onto the deck[8]. For typical geometries, these
loadings cause deflections in the central plating that are
so large when compared with the plate thickness that
bending is negligible and the plating is analyzed and
designed as a membrane. The radial membrane forces
are reacted by the pontoon ring. The ring is pulled
inward and, therefore, compressed tangentially. The ring
is also twisted and, hence, is subjected to bending stresses normal to the cross section[9]. The local and overall
stability of the pontoon ring have aiso been
considered [lo].
The complicated nature of the loading on the membrane together with the unusual boundary conditions
requires numerical integration of the coupled, non-linear
differential equations which govern the behavior. A
skilled analyst is needed to vary the starting parameters

349

HOWARD
I. EPSTEIN

350

necessary for the integration. This paper desoribes the


procedure that should be followed in the analysis and
design of these roofs, and shows why an interactive
computer is so important in the understanding of the
forces present and in the design necessary to transmit these
forces.

(3)
where

EQUILIBRIUM
EQUATIONS

basic parameters for the membrane/pontoon


problem are illustrated in Fig. 1. The overall geometry
with the membrane filled with rainwater is shown in Fig.
l(a). Values for HO,H, and w, are unknown until the
final equilibrium position for this system is determined.
Note that a negative value for H, is possible as this
would represent a partially filled membrane. The membrane is assumed to be pinned to the pontoon at point E.
The membrane force is transmitted to the pontoon at
point E causing a horizontal displacement, 8, and a
rotation, I$.
The

Membrane equations
The membrane geometry is shown in Fig. l(b). At any
radius, r, the vertical downward displacement relative to
the edge is defined as w, and the outward radial displacement as u. The radial stress in the membrane at the center
(r = 0) is defined as a, and at the edge (r = d) the stress is
defined as crc.The rotation of the membrane edge is given
by a and the thickness of the membrane is t. The lateral
pressure applied to the membrane at any point below the
water surface is given by
P = (w + H&L -(w + H&x,

(1)

where p,,, and p0 are the rainwater and oil densities,


respectively, and where a positive pressure is downward.
Equilibrium of an element of the membrane leads to two
coupled, first-order, nonlinear, differential equations in
terms of the vertical displacement, w, and the radial
force per unit length, N,. Mitchell[lll presented the
differential equations for the large deflection of a circular, symmetrically loaded membrane[l2,
131 by
assuming bending terms to be negligibly small. These
equations are

Fig. 1. Geometryof membrane/pontoonconfiguration.

and where E = Youngs modulus.


Solutions to the membrane equations may be accomplished by using the Runge-Kutta
integration
method[7, II]. The equations are integrated from the
center (i = 0) to the edge (P = 1) in small increments, A?.
In order to start the integration, the center tension, a,,
the depth of water at the center, 0, and the oil head, AH,
must all be spectied. Basically, the overall solution to
the problem is accomplished by systematically varying
these parameters until all the conditions imposed on the
problem are satisfied.
The radial displacement of the membrane, u, is not
necessary in the integration procedure. However, it is
necessary to keep track of this displacement in order to
establish the horizontal movement of the edge of the
membrane, 8. The resulting equation to find 8 is[7]
6=tj)[;;($!~-(1-vz)&]di

(5)

where v = Poissons ratio. This equation is solved


numerically as the integration of eqns (1) and (2) is being
accomplished.
The volume of water contained in the membrane, Q,
may also be found numerically as the integration is
proceeding. For a full membrane, this volume is given
by PI
Q=2?rd2

[Dtw-w.)idr
I0

(6)

The upper limit in the integration changes when the


membrane is only partially full.
Pontoon equations
The pontoon geometry is shown in Fig. l(c). A hollow
trapezoidal cross section of uniform thickness, g, is used
in this presentation. The interior of the ring-shaped pontoon is divided into a number of compartments by solid
radial ribs. The cross-sectional dimensions of the pontoon are usually chosen so that buoyancy is insured even
with two compartments ruptured[l4], and so that the
design rainfall can be contained. These restraints can be
satisfied with a variety of cross section con6gurations.
However, large vertical dimensions for the pontoon
would require excessive freeboard allowance. Also, it is
desirable to slope the bottom of the pontoon to diiect
any vapors formed under the roof away from the rim
space and to slope the top of the pontoon to direct any
rainfall toward the center. All these considerations
naturally lead to a wide, rather shallow trapezoidal cross
section.
If the pontoon is assumed not to rotate when loaded
only with its own weight, the center of rotation corresponds to the center of gravity (point 0 in the figure)
and is located at a radius R from the center of the
membrane. When the radial membrane force is applied to
the pontoon, the tendency to rotate is resisted by the oil
pressure along the bottom and the restoring moment in

Floating roof analysis and design using minicomputers

the pontoon itself. Eliminating the rotation from the


vertical and rotational equilibrium equations leads to
17, 111
qc =

UJ C, sin a

t G cos a) + Csu.

sin a cos a

c, + c5f.7.cos a

(7)

where q. is the oil pressure at point E and where C, to


CJ are functions of the pontoon geometry, the material
properties of the pontoon and the oil density.
The horizontal component of the radial membrane
tensile force per unit length is given by o,f cos a. This
produces an inward displacement of the pontoon
S=
where

u.tR2 cos a
AE

(8)

A = the cross-sectional area of the pontoon.

0.2

a4

0.6

0.6

1.0

Possible equilibrium positions


Values are required for the oil head, center water
depth, and center tension in order to start the numerical
integration in the membrane. When the step-by-step integration for particular starting values is carried to the
edge of the membrane, o, a and q. are determined.
There are three conditions that must be satisfied to
insure that the solution thus obtained is the correct one:
First, the horizontal movement of the edge of the membrane numerically found from eqn (5), must equal that of
the pontoon as calculated from eqn (8). Second, the
water contained in the membrane, as calculated from eqn
(6), must be equal to a predetermined amount, and; third,
the edge quantities, cr., a, and q. found by the integration must satisfy eqn (7). Since the edge stress is
contained in eqn (8), the pontoon is providing a horizontal stiffness to the edge of the membrane. If a stifYness,k,
is defined as the force per unit circumferential length
required for a unit radial displacement, eqn (8) gives
k=F

AE

Thus, the matching of eqns (5) and (8) is assuring


horizontal compatibility, while the matching of q* as
given by eqn (7) is assuring vertical compatibility at the
interface.
Effect of varying the initial parameters
Each of the three parameters for starting the integration has its own effect on the final shape of the
membrane and the position, slope and stress at the edge.
It has been shown that a systematic variation of these
three parameters can lead to the final equilibrium
position[7]. For given values of D and AH, the higher the
value for k, as given by eqn (9), the larger is the required
center tension. Therefore, different horizontal stiffnesses
can be matched by changing the center tension and
hence horizontal compatibility can be assured.
The effects of varying D and AH can be seen in Figs.
2 and 3. It is fist assumed that AH is given and D is
allowd to vary. Typical results are shown in Fig. 2 where
the deflected shape of the membrane is plotted for a
given AH and varying central water depths, D, > DZ>
. . . > 4. For each depth, the central tension is adjusted
to give a specific stillness at the edge. Ds (referred to as
the floating depth) is the depth at which the oil pressure

Fig. 2. Typicaldeflected shapes for the membrane.

Fig. 3. Effects of changes in membrane parameters.


from below equals the water pressure from above. Below
that depth, the net pressure is downward and therefore
the curvature of the membrane is upward. The reverse is
true above that depth, and, therefore, when the central
depth is close to the floating depth, the deflected shape of
the membrane tends to oscillate about the floating depth.
For ranges of central depth, a plot showing the edge oil
pressure, q.. vs edge rotation, a, is given in Fig. 3(a).
Large central depths give large q* and a. As the central
depth decreases to the floating depth, the curve spirals
inward to the crossed point representing the oil pressure
at the floating level (zero edge rotation). A further
decreasing of the depth produces another branch of the
curve which spirals outward until zero central depth is
reached.
If the oil head. AH, is increased, the floating depth is
also increased, and the plot in Fig. 3(a) is shifted upward
by a uniform amount (as partially shown by the upper
dashed curve). Similarly, decreasing the oil head lowers
the curve, except that sufficiently decreasing the oil head
results in cases of partially filled membranes (the membrane extends above the water surface) and this portion
of the curve cannot be obtained by simply shifting the
plot downward. As D increases further, some of the
partially filled cases result in membranes which project
above the oil surfacl (qe CO). This condition is physically not possible.
For a given AH, as D varies, a q. vs a spiral is
generated as shown in Fig. 3(a). The volume of contained
water, Q, also changes as D varies, as can be seen in Fig.
2. Plots of constant Q curves are shown in Fig. 3(b). For
a particular volume Q, each point along the correspond-

352

HOWARD
I. EPSTEIN

ing curve represents a possible equilibrium con&ration.


If the center tension has been adjusted to satisfy
horizontal compatibility with the pontoon, the vertical
and force compatibility and, hence, the unique equilibrium position can be found by determining the
configuration which also satisfies eqn (7).
SOLUTIONPROCEDURE

A flow chart of the overall process which is used in the


analysis and design of floating roofs is shown in Fig. 4.
Seven basic steps are identified in the process.
The input necessary in Step 1 is: membrane and pontoon geometry and material properties; the densities of
the supporting and contained liquids, and; the conditions
for which the system must be designed. Step 2 selects
the starting parameters necessary for the numerical integration. Step 3 accomplishes the numerical integration
of eqns (2) and (3) as well as (5) and (6).
If, in Step 4, eqns (5) and (8) give identical results,
horizontal compatibility is assured, and the solution can
proceed; if not, the center tension is adjusted until the
numbers converge. At this point it should be noted that
some combinations of D and AH do not result in possible equilibrium positions[7] and for these, the process
must be restarted.
Once Step 4 is passed, one point on a constant AH
spiral (see Fig. 3a) has been found. In Step 5, if the
volume of contained liquid, as calculated from eqn (6), is
not what it should be, more of the spiral is generated
until one point on the correct constant Q curve (see Fig.
3b) is generated. If this point also satisfies eqn (7) in Step
6, vertical and force compatibility is assured, and the
unique equilibrium position has been found; if not, a new
AH is assumed and the process is repeated.
The final equilibrium position obtained by this process
can be used in finding the forces and stresses in the
membrane and pontoon[7,11]. It has been shown that
typical floating roof pontoons are subjected to considerable axial stresses considering the geometry of the cross

rei
eff

t
Fig. 4. Flow chart.

ocai@~
O.K.

section. This introduces a final complication into the


process of analysis and design of pontoon floating roofs
and that is the presence of local buckling of the pontoon
cross section as an active factor. If local buckling is a
consideration, only a fraction of the cross-sectional area
can be considered to be effective. This area is a function
of the axial forces on the pontoon cross section which, in
turn, is a function of the effective area.
In Step 7, the equilibrium position found in Step 6
must be examined in light of the stresses found and the
local and overall buckling criteria. If the design is inadequate, or the effective area calculated is not in accordance with that which was assumed in finding the
equilibrium position and associated forces, the process
must be repeated. The effect of local and overall buckling of the pontoon and the iterative solution procedure
to be followed when these factors are present have been
investigated [IO].
THEUSEOF AN INTJ%RACTIVE
COMfUTER
is possible to write a computer program to accomplish all the steps shown in the flow chart in Fig. 4.
However, there are many difficulties which must be
resolved in order to write such a program including, but
not limited to:
0 In which directions should the starting parameters
be incremented?
0 What step sixes should be used?
l How and when should these step sixes be changed?
0 How to recognize inadmissible starting parameters.
0 What to do when the design is inadequate.
0 What constitutes a good design?
There are several reasons why the writing of such a
program has not been undertaken and these deal mainly
with the last two areas mentioned above.
The problem is ideally suited for solution on an interactive computer. Once the effects of varying the starting
parameters of the integration (as shown in Fig. 3) are
understood, finding an equilibrium position for any given
input data (including effective area) can be accomplished
routinely. After a little experience, efficient step sixes
and directions, which are needed to quickly solve the
problem, can usually be found.
In going through the steps necessary to find an equilibrium position, the interactive computer gives insight
into the effect of the starting parameters, the displaced
conIiguration of the membrane and pontoon, and the way
in which the internal forces are transmitted. This insight
enables the engineer to thoughtfully answer questions
such as:
0 What is the importance of the location of the
attachment point between the membrane and the pontoon?
0 Is it necessary to consider intermediate equilibrium
states, or just the final position, when the decking is
being filled with water or the product is leaking through
punctures onto the deck?
0 Is the so-called sag-full condition (Hw = 0) a
reasonable water level to use as a design condition?
0 What effect will changing the geometry of the pontoon have on the stresses and displacements?
The engineer becomes an integral part of the solution
process by determining what should be changed if the
design proves inadequate and in estimating what the
effects of these changes will be in the next analysis
sequence. It is the interaction with the analysis process
that gives the engineer the feel for the structure. This
It

Floating roof analysis and design using minicomputers


is a very important part of the process and it is the main

reason that an interactive solution procedure is so vital in


solving the problem.
Various pontoon floating roofs have been analyzed on
an interactive time-sharing terminal and on a desk-top
minicomputer. It doesnt make much difference what
system is used as long as it possesses an interactive
mode and solves each iteration in a reasonable time
period since several hundred iterations may be necessary
to analyze the roof.
REFERENCES

I. 0. 0. Storaasli, On the role of minicomputers in structural


design. Comoul. Strucfures 7. 117-123(1977).
2. J. A: Swans& Use of mini-computersfor iarge scale structural analysis programs. Comput. Structures 7, 291-294
(1977).
3. II. W. Haeseker and C. S. Hodge, West coast consulting firm
gets larger minicomputer to keep pace with growth. Civil
Engineering-ASCE, Apr. 60-65 (1978).
4. J. de Wit, Floating roof tanks. Engng 210,55-58 (July 1970).
5. The storage of volatile liquids. Tech. Bull. No. 20. Chicago
Bridge and Iron Co., Plainfield, Illinois (1947).

CAS Vol. II. No. 4-C

353

6. W. B. Young, Floating roofs-their design and application.


ASME Petroleum Mech. Engng Con/. Los Angeles, California (16-20 Sept. 1973)(Preprint 73-Pet-44).
7. H. I. Epstein and J. R. Buzek, Stresses in floating roofs. J.
Srrucr. Div.. ASCE 104.735-748 (1978).
8. H. 1. Epstein and J. R.Buzek, Sires& in ruptured floating
roofs. .I. Pressure Vessel Technology, Trans. ASME 100,
29l-2% (1978).
9. J. F. Harvey, Pressure Vessel Design. Van Nostrand, Princeton, New Jersey (1%3).
10. H. I. Epstein and J. R. Buzek, Design of pontoons for
floating roofs. ASMEICSME Pressure Vessels and Piping
Conf. Montreal, Canada, (25-30 June 1978(Preprint 78-PVPIll).
11. G. C. Mitchell, Analysis and stability of floating roofs. J.
Engng Mech. Div., ASCE 99, 1037-1052(1973).
12. E. H. Mansfield, The Bending and Stretching of Plates.
MacMillan, New York (1964).
13. T. von Karman, Festigkeitsprobleme in mashinenbrau.
Encyklopadie der mathematishen Wissenschaften IV, 349
(1910).
14. Welded Steel Tanks for Oil Storage (5th Edn.). American
Petroleum Institute Standard 650 (1973).

You might also like