Sei sulla pagina 1di 17

THE STRUCTURAL DESIGN OF TALL AND SPECIAL BUILDINGS Struct. Design Tall Spec. Build.

16, 501517 (2007) Published online 2 November 2007 in Wiley Interscience (www.interscience.wiley.com). DOI: 10.1002/tal.413

THE DAMPED OUTRIGGER CONCEPT FOR TALL BUILDINGS


ROB J. SMITH AND MICHAEL R. WILLFORD*
Arup, London, UK

SUMMARY This paper describes new concepts for the structural design of high-rise buildings, in which a system is introduced to increase the dependable structural damping by a factor of 510. By so doing the dynamic response of the building to wind effects (buffeting and vortex shedding) is virtually eliminated, leading to substantially reduced lateral design forces and assured occupant comfort. Substantial reductions in structural member size and construction cost savings can be realized in many cases. This may signicantly improve the economic viability and sustainability of a development. The paper describes some means by which high levels of damping may be achieved and is illustrated by an implementation on a building, currently under construction, subjected to wind and seismic excitation. Copyright 2007 John Wiley & Sons, Ltd.

1. 1.1 Dynamic response of tall buildings

INTRODUCTION

It is often found, particularly for very slender buildings, that the dynamic resonant response of the structure to incident wind gusts and from vortex shedding leads to a signicant increase in the strength that has to be provided in the lateral resisting structure. It is also common for the predicted wind-induced sway motion of a building to be excessive for human comfort, and this may require the dynamic response to be reduced. In seismic regions, earthquakes also induce strong lateral motions, and reduction of the lateral dynamic response is desirable in this case to reduce damage levels. The dynamic response of a tall building is governed by a number of factors including shape, stiffness, mass and the damping. While the effect of shape can be assessed by wind tunnel testing, and the mass and stiffness can be predicted with reasonable accuracy by the structural designer, the intrinsic damping of tall buildings is low and the guidance available shows huge variance. By adding an engineered supplementary damping system to a building, it is possible to remove dependence on the low and uncertain intrinsic damping. This improves the reliability of dynamic response predictions and, by supplying higher levels of damping, substantially reduces the required stiffness of the building while at the same time improving performance. This paper describes methods by which high levels of dependable damping may be introduced into high-rise building structures, and the benets that may be realized.

*Correspondence to: Michael R Willford, Arup, 13 Fitzroy Street, London W1T 4BQ, UK. E-mail: michael.willford@arup. com

Copyright 2007 John Wiley & Sons, Ltd.

502 1.2

R. J. SMITH AND M. R. WILLFORD

Intrinsic damping in tall buildings

Data on the effective damping in high-rise buildings can only be obtained through measurements on existing buildings. The best collation of high-quality data on modern high-rise buildings is by Satake et al. (2003). We present these data, alongside other recent measurements (Jeary, Li, Brownjohn, Wood and others) in Figure 1. It can be seen that there is very considerable scatter in the measured data. It is clear, though, that the trend is for intrinsic damping to become lower and less variable as building height increases. There are logical reasons for this, associated with the relative signicance of energy dissipation in cladding and t-out as buildings become taller and more slender. While precise formulae cannot be tted to the data, it is evident that it is most likely that damping will be between 05% and 1% for buildings above 250 m in height. While it is generally understood that damping increases with response amplitude, measurements by Tamura (2006) show that this is not always the case, and that the highest damping can occur at relatively small response levels. This is consistent with expectation if a proportion of the damping is provided by frictional resistance in cladding, for example. Figure 2 shows the variation of measured damping with response level for three buildings.

Figure 1. Measured intrinsic damping of tall buildings

Figure 2. Examples of measured damping versus response level


Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

THE DAMPED OUTRIGGER CONCEPT FOR TALL BUILDINGS

503

The uncertainty and variability in the data present risks for building design. It is advisable to select a conservative value for the damping and to design the structure accordingly, but this in not always done. Many practitioners and wind tunnel laboratories recommend damping levels in the range 12% of critical, and while this might be appropriate for many buildings lower than 200 m, it could be a factor of two too high for taller buildings. While the main objective of adding damping via specially engineered components is to reduce building motion and in some cases design forces, a side effect is that the uncertainty in damping level can also be reduced. 1.3 Mitigation of dynamic response Increase in stiffness Conventionally, the design approach taken to control dynamic wind-induced response is to increase the stiffness and strength of the lateral resisting structure. Any increase in stiffness reduces the natural period of the building, which generally leads to a reduction in dynamic response. The strength of the lateral resisting system is then made sufcient to resist the dynamically enhanced wind load effects at the lower natural period. The provision of increased stiffness and strength leads to larger structural member sizes relative to those required if there were lower resonant dynamic response. In addition, for seismic regions, increasing the stiffness increases the seismic loading, owing to the stronger ground motions associated with shorter periods. Tuned mass/liquid dampers If the predicted building motions are excessive for occupant comfort after strength and stiffness criteria have been satised, a tuned mass or tuned liquid damper at the top of the building is often specied. While a tuned mass/liquid damper system can often be designed to provide the equivalent of 24% of critical damping (which might halve the resonant response), these devices have a number of disadvantages:

they are large, heavy, and take up valuable space at the top of the building; they have to be tuned closely to the measured natural frequency of the building mode of concern if there are several modes of concern then several sets of differently tuned devices are required; they are an additional cost to a project.
There is reluctance to rely upon tuned mass/liquid devices to reduce the strength design forces on a building because: the natural period of the building might change with time and with response amplitude such that the device becomes de-tuned and less effective; it is more economical in terms of cost and space to provide one large damper unit. There is therefore a question of dependability. If the unit or one component of it were to fail, there would be no backup; it is not practical to design these devices to provide control under severe seismic events. For these reasons they are only used to reduce occupant perception of motion. Active mass damper devices enable greater control to be achieved with a given mass than with passive devices, but are more complex, require powered components and computer control, and are considerably more expensive than passive devices. In addition, their reliability under extreme loadings is uncertain, particularly as they require a power supply.

1.4 Viscous/viscoelastic dampers and similar devices An alternative means of providing damping is to introduce resistance devices such as viscous dampers, viscoelastic dampers or active stiffness control devices. These devices are attached between
Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

504

R. J. SMITH AND M. R. WILLFORD

two points on a structure and operate on the relative motion between those two points. They are best located to connect two points having signicant relative displacement in the mode of concern. These devices do not require frequency tuning, and while there is an optimum resistance characteristic for each application, the overall damping achieved is usually not very sensitive to the exact resistance characteristic of the device. It is therefore sometimes possible to damp several modes with one device, provided the two points it connects have signicant relative movement in each of the several vibrating modes of the structure. A number of generic types are available, all of which operate upon the relative motion between their ends. Such devices have been used to reduce seismic damage in bridges and buildings, and also to reduce vibrations on large machine foundations. Components may be categorized under the following headings:

Viscoelastic dampers Fluid viscous dampers Viscous pot and plate dampers Friction dampers
Viscoelastic dampers are devices in which layers of special high-damping solid polymers are deformed in shear under the cyclic motions between the two ends of the device. An early example of their use in building structures was in the World Trade Center towers in New York. These devices respond linearly at small strains, although the stiffness and damping properties of these materials are frequency and temperature sensitive. Fluid viscous dampers are piston-type devices with arrangements of seals and piston orices to provide a resistance force as a function of velocity between the two ends of the device. Different manufacturers supply a wide variety of devices, operating at high or low uid pressure and generally having non-linear force versus velocity relationships of the form F = Cva, with a in the range 0420. These devices have previously been used for seismic applications (Higashino and Okamoto, 2006) and in bracing arrangements for retrotting buildings (McNamara et al., 2003). Viscous pot and plate dampers are similar to viscoelastic devices but using high-viscosity uids in shear rather than solid viscoelastics. Friction dampers rely upon the friction between two suitable surfaces having a compressive force maintained between them. Energy is dissipated in sliding frictional resistance once the static friction limit has been overcome.

2. 2.1

DESCRIPTION OF THE DAMPED OUTRIGGER AND VARIANTS

Review of structural forms for tall buildings

Tall buildings are constructed incorporating one or more structural systems to provide resistance to lateral forces. Typical systems for high-rise buildings include:

core connected to external columns with outriggers (and belt trusses); coupled cross-walls; external tube structure.
Different forms have different applications depending on architecture and building use. The current paper discusses novel variants of these structural systems, in which viscous damper elements are introduced into the load paths in order to generate dynamic stiffness in the form of damping resistance, at the expense of static stiffness and strength. If the damping generated is
Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

THE DAMPED OUTRIGGER CONCEPT FOR TALL BUILDINGS

505

sufcient, the reduction in dynamic lateral response (lateral force) more than compensates for the reduction of static stiffness and strength, resulting in a more economic design. 2.2 Overview of damped outrigger and variants

The concept is illustrated in Figures 36. Figure 3 shows how the system works in simple terms when embodied within a conventional core-to-perimeter column outrigger system. As a building undergoes dynamic sway motion, there is relative vertical motion between the perimeter columns and the ends of stiff outrigger elements cantilevering from the core. A damper is inserted across this structural discontinuity, dissipating energy during the cyclic motion, and resulting in an increase in the overall damping of the building. Figure 4 shows an example of another arrangement. In this case a conventional coupled shear wall is adapted by making the normal coupling beams as exible as possible, and introducing a small number of deeper corbel outriggers at various levels connected by vertically acting dampers. Other arrangements are possible. Figure 5 shows conceptually the form of detail typically required at the levels at which dampers are incorporated. It is necessary for the outrigger element to move vertically relative to the oors at these levels, while the oors bend in double curvature to remain connected to the core and the outer columns. This can be acceptable when outriggers are provided in plant room levels, provided items crossing the outrigger are appropriately supported. If outriggers are provided at occupied levels it is necessary to provide architectural stud walls independent of the outrigger supported on the oors to each side of the outrigger itself. In Figure 5, two separate damping components are installed at the end of the wall. It is economically possible to supply the damping resistance in the form of a number of smaller components and this increases the redundancy of the system. The arrangement at an entire outrigger level in this case could be as shown in Figure 6.

Figure 3. Damped outrigger concept 1 (patent pending)


Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

506

R. J. SMITH AND M. R. WILLFORD

Short outriggers /corbels Vertically acting dampers

Shear walls

Figure 4. Damping concept 2: damped shear wall (patent pending)

Figure 5. Conceptual detail at outrigger level

2.3

Performance of these damping systems

The performance of this type of system depends on a number of factors including:

the exural and shear stiffness of the various core or wall elements; the axial stiffness of the perimeter columns and their distance from the core;
Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

THE DAMPED OUTRIGGER CONCEPT FOR TALL BUILDINGS

507

Figure 6. Typical layout at outrigger levels

the number, position and stiffness of the outriggers; the stiffness of oor beams spanning between the core and the perimeter, or coupling two halves of a cross-wall or core, at typical levels; the stiffness of other elements of the lateral resisting system of the building, e.g. perimeter frame
action. In many tall building structures in is possible to achieve a damping value of between 2% and 15% of critical using these techniques. In practical terms, a damping ratio of 810% is the maximum that would usually be required to virtually eliminate the resonant enhancement of wind load. 2.4 Calculation of overall building damping In general, for practical design, it is necessary to calculate the effect of a proposed damping system and the optimum resistance of the damper elements by direct solution dynamic nite element analysis of a model of the proposed building. This is because the structural congurations vary from building to building and from level to level within a building. Direct solutions are dened here as solutions which solve the dynamic equations of motion directly from the mass, stiffness and damping matrices. These include complex modal analysis and direct harmonic forced response analysis. Methods involving normal modes are unsuitable because they cannot account for discrete damping elements within a structure. Figure 7 shows the output from a steady-state forced response analysis with a constant lateral harmonic force of varying frequency. The damping can be estimated by the half power bandwidth method, as would be the case with a transfer function of this type obtained experimentally in a real building. 3. COMPARISON WITH CONVENTIONAL DESIGN APPROACHES 3.1 Designing without supplementary damping Conventional design of tall buildings generally excludes the use of supplementary damping systems. For low- to medium-rise buildings damping has little effect on the behaviour of the building, but for
Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

508

R. J. SMITH AND M. R. WILLFORD

Figure 7. Typical response transfer function

taller buildings resonant dynamic response under wind can become very signicant and will lead to unacceptable motions for occupant comfort. Motions are conventionally controlled by stiffening the structure. The effect can be seen graphically in Figure 8 obtained by processing wind tunnel high-frequency force balance measurements for a 400 m high tower. Here, overturning moment and acceleration are plotted against building natural period for different levels of damping. For a building without supplementary damping, intrinsic damping of 08% has been assumed, consistent with Figure 1. In this case, the natural period needs to be reduced to approximately 7 s to meet the ISO acceleration criteria. The corresponding overturning moment is approximately 23 000 MN m. 3.2 Designing with tuned-mass dampers Where lateral accelerations exceed occupant comfort criteria a tuned-mass damper is often incorporated to control motions. Typically, this may provide a damping equivalent to approximately 24%, giving a total damping of 35%. As can be seen in Figure 8, the lateral accelerations are reduced to acceptable levels for a range of natural periods. This might enable a more exible structure to be designed incorporating less material. However, as explained earlier in this document, it is not possible to use a TMD to reduce design loads. Therefore the design for strength must assume no supplementary damping. Since the consequence of making a building more exible without allowing for increased damping is to increase the design loading, there is limited opportunity actually to remove material from the structure. 3.3 Designing using viscously damped outriggers

Using highly dependable and redundant viscous damper arrangements, it is possible to reduce both lateral accelerations and design loads. In the example below, if it is assumed that the total damping in the structure can be increased to 5% of critical, then the overturning moment used for design is approximately 12 000 MN m, i.e., half that of the conventional design. In addition, the range of acceptable stiffness is large. This represents signicant savings in structural material.
Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

THE DAMPED OUTRIGGER CONCEPT FOR TALL BUILDINGS


= 0.8 % = 2.0 %

509

Overturning Moment (MNm)

30000

= 5.0 %

20000
X

10000

0 6 7 8 9 10 Period of Oscillation (s) 11 12

= 0.8 %

Peak Acceleration (milli-g)

40 30 20 10 0 6 7 8 9 10 Period of Oscillation (s) X T V

= 2.0 % = 5.0 % ISO Criteria - limit

11

12

Figure 8. Sensitivity of dynamic response for different damping systems applied to 400 m high tower. Results are taken from a wind tunnel test. Lateral accelerations are for the top of the building during a 10-year storm. Location is in a typhoon region. X, undamped; T, tuned mass damper; V, viscous damper

However, in order to mitigate against the risk of catastrophic damper failure, in this case we would recommend designing for a slightly higher overturning moment of 14 000 MN m. See later discussion on reliability. 4. DESIGN PROCEDURE

There are two sets of tasks which need to be performed to design a damping system. 4.1 Choose required stiffness and damping level for the building

For a building of a given size, shape and mass, the dynamic performance is governed by the stiffness and damping. By varying these, it is possible to plot the dependence of structural loads and lateral accelerations on these properties. These predictions are ideally made using either wind tunnel test data (typically results from a high-frequency force balance test), or from an empirical method such as that
Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

510

R. J. SMITH AND M. R. WILLFORD

in the National Building Code of Canada (NBCC). Typical sensitivity graphs are shown in Figure 8. Here stiffness is characterized by natural period. 4.2 Design damping system, estimate damping and optimize structural stiffness

Viscous dampers are ideally located between elements having high stiffness and which, unconnected, experience high relative movement. Suitable locations are typically at the ends of outriggers, between shear walls and in diagonal braces. Note that the latter case is usually suited to frame buildings. Damping is calculated using the direct solution nite element analysis methods described previously. It is necessary to iterate the viscous resistance (C value) of the dampers to determine its optimum value. The damper locations and resistances and the global structural stiffness can then be optimized either by manual trial and error or by automated techniques. Once this is done, the structural loading based on the earlier sensitivity analysis can be reassessed. Strength and deection checks then need to be reperformed. 4.3 Note regarding nature of damping force

The dynamic (resonant) uctuating force in a viscous damper is 90 degrees out of phase with the elastic force in conventional structural elements. For that reason, the combination of the resonant element forces and elastic forces is made thus: Total_response = Mean Gust 2 + Dynamic 2 Quasi-static = Mean Gust The resonant forces should be calculated by frequency response analysis or similar methods scaled to the peak resonant response level under the enhanced damping. The mean and gust force distribution can be calculated by static analysis. 5. 5.1 PRACTICAL DESIGN CONSIDERATIONS

Power dissipation and cooling

The action of the dampers results in heat generation in the damping units. While the heat generated in 30 seconds of strong earthquake may be absorbed in the thermal inertia of the units, for long-duration wind storms a steady-state temperature will be reached and the heat generated must be dissipated from the damper bodies. Typical heat generation is in the region of 13 kW per damper. Since this heat is only generated during wind storms, passive ventilation created by wind can generally be used to cool the units. 5.2 Fatigue and durability

The dampers and their connections are subjected to many millions of cycles of wind-induced forces over a typical building lifetime. It is therefore necessary to check the fatigue life of the most critical components. This includes connections, seals, pistons, bearings, etc. The fatigue life of the dampers may be enhanced by increasing the velocity exponent (a in the equation F = Cva of the damper. In doing this, the damper is subject to lower forces for regularly occurring winds. In doing this, the effectiveness of the damper is reduced at lower wind speeds. Consideration should be given to lateral accelerations at lower wind speeds.
Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

THE DAMPED OUTRIGGER CONCEPT FOR TALL BUILDINGS

511

5.3 Reliability, maintenance, replacement and access Although the dampers have a high level of reliability and spare capacity within the system, they are still required to be inspected and maintained to ensure adequate long-term performance. Inspection should be carried out under an agreed schedule to examine the following:

checking for leaks; appearance of cracks; signs of metal fatigue; corrosion.


Should any problems occur, then remedial action should be taken according to the suppliers manual. Inspection should also be carried out after an earthquake or major wind storm. It is expected that the dampers will need to be replaced after approximately 50 years, so future access needs to be designed into the system. These requirements mean that a building owner must be informed of his responsibility to monitor and maintain the dampers, just as he would for all aspects of the building, such as facade, lifts, water supply, etc. 5.4 Robustness

As long as dampers are supplied by a reputable manufacturer and are routinely inspected and maintained, they are extremely unlikely to fail during their lifetime. However, this possibility must be anticipated and any unacceptable consequences mitigated. By providing a number of separate damper units to provide the overall damping resistance the effect of the possible failure of a proportion of the units can be controlled. Figure 9 shows the dependence of the overall overturning moment on a building with the overall viscous resistance of the damper system. It can be seen that there is an optimum resistance, but that providing a resistance somewhat greater than optimum does not increase the moment signicantly. By providing more than the optimum overall viscous resistance through a signicant number of individual units, it can be seen that failure of some of those units (reducing the total resistance) will have little overall effect on the buildings performance.

6
Overturning Moment [GNm]

Design value

0
0 1000 2000 3000 4000 5000
C coefficient of dampers [MNs/m]

Figure 9. Variation of overturning moment with overall viscous resistance


Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

512 5.5 Building deection

R. J. SMITH AND M. R. WILLFORD

For tall buildings without supplementary damping, the governing criterion for lateral stiffness is usually occupant comfort under lateral acceleration. For buildings with high supplementary damping, lateral accelerations are acceptable and the next governing criterion can become lateral deection. Traditional deection limits such as H/500 are to a large extent arbitrary, and need not apply to highly damped buildings. Guidance given in national design codes varies signicantly and in many cases higher deection can be accommodated by appropriate detailing of facades, partitions, etc. What the real functional limits on lateral deection of modern high-rise buildings are is an area of ongoing research. 5.6 Fire and weather protection

Dampers should be protected from the effects of re and weather, in the same way that conventional structural components would be. 5.7 Differential vertical shortening

Differential shortening between a core and the exterior columns of a high-rise building is a common concern when outriggers are employed, particularly for reinforced concrete structures. Connecting the columns to the outriggers by means of dampers means that static differential shortening can occur without generating any force in the outriggers. The design damper stroke must make allowance for the expected amount of differential shortening in addition to performance under lateral effects. 5.8 Use for seismic design

Although these damping systems are principally intended to reduce dynamic wind-induced response, they are also suitable for reducing seismic response. Benet from the incorporation of viscous dampers can only be evaluated by means of appropriate non-linear response history analyses in the context of a performance-based seismic design approach. The authors would always recommend such an approach for a tall building, where higher modes dominate the behaviour, static methods are unsuitable and conventional code-based design based on ductility-modied response spectrum analysis can be both unsafe and uneconomical. It should be noted that the displacement of the dampers (the stroke) and the relative velocities can be considerably higher in a seismic event than a windstorm. For this reason, a force-limiting device in the damper will often be necessary. Different suppliers have different methods of providing this, but a pressure release valve is one option. 5.9 Variation of velocity exponent Damper manufacturers usually provide dampers where the force generated, F, is related to the relative velocity between the ends by F = Cv where a is between 01 and 2. This is done by altering the shapes and sizes of the various orices within the damper cavity. The choice of exponent depends on a number of factors:
Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

THE DAMPED OUTRIGGER CONCEPT FOR TALL BUILDINGS

513

the technology of the manufacturer; whether high levels of damping are required at low or high amplitude or both.
5.10 Risks involved in using dampers to reduce design loads Since the dampers require maintenance (albeit very little), there exists a residual risk that malfunction of the damping system may cause structural failure. Given the redundancy associated with providing many individual damping units and the high integrity of the dampers, the risks that could be considered are:

abandonment of the building during construction before dampers are installed; abandonment of building once complete (i.e., no building owner to maintain the building); neglect of dampers during building occupation; systematic failure/poor performance of dampers (i.e., manufacturing problem); sabotage.
Typical mitigation measures could be:

Limit the degree to which dampers are used to reduce the moments for strength design to M

damped > rMundamped, where M = overturning moment and r is a reduction factor. This limitation ensures that in the case of total damper failure the building will remain standing in the design wind. The reduction factor should be considered on a case-by-case basis, but should consider the likely material overstrength, the acceptable reduction in load factor and suitable intrinsic damping ratios. The likely value is in the region of r = 0507. This limitation need not be applied to deection checks. Ensure that dampers are of high integrity with minimal servicing.

6.

DESIGN EXAMPLE

This example is a development of two similar residential buildings 210 m tall and approximately 38 m square in plan located in a region of typhoon winds and UBC Zone 4 seismic conditions. Each building has a reinforced concrete core (coupled in one direction) and an irregular arrangement of perimeter columns and walls. One level of double storey height outriggers is present at approximately half the overall height. This section considers the load effects in one direction of one tower only. At the time of writing this building is under construction, with the viscous dampers manufactured and awaiting test. 6.1 Seismic design

The building is in UBC 97 Zone 4 location with a nearby fault. The seismic design has used performance-based design methods, using modelling and deformation acceptance criteria based on FEMA 356. The performance-based (deformation-based) approach results in seismic loading no longer governing much of the structural design. For the purposes of brevity, the seismic design of this development is not discussed further here. 6.2 Wind effects

The building is subjected to typhoon winds. The likely dynamic response was initially calculated using the Detailed Method of the National Building Code of Canada for both along-wind and across-wind directions. For occupant comfort 10-year and 1-year return periods were considered, and 100-year
Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

514

R. J. SMITH AND M. R. WILLFORD

return period was taken for strength design. Following initial design, wind tunnel testing was used to better quantify the wind loading. Alongside the wind tunnel tests, a directional climate study was performed. This made a signicant difference to the nal assessment of wind-induced response. Considering one load direction only, the base overturning moments were as shown in Table 1. Two observations can be made from comparing this table with the code loading table:

The undamped wind overturning moment predicted by wind tunnel testing is more than the ASCE prediction but less than the NBCC. The use of dampers alongside a climate study has reduced the overturning moment by 50%, representing a signicant reduction in load and associated structural cost. Lateral accelerations 6.3

The lateral accelerations shown in Table 2 were predicted for the 10-year wind, using wind tunnel tests. From these results it is clear that the addition of damping substantially reduces lateral accelerations below the recommended limit for the return period. 6.4 Implementation of damping system Layout For each of the two buildings, eight outrigger walls are attached to the core approximately halfway up the building. Two dampers are attached to the end of each of these outrigger walls, i.e., a total of 16 dampers per building. The dampers have been placed side by side for ease of construction and because of restrictions on space. This is shown in Figures 5 and 6.

Table 1. Variation of wind loading with damping Effect Wind Wind Wind Wind Wind Wind Method NBCC NSCP/ASCE NBCC WTT WTT WTT Assumption Code wind speed, intrinsic damping (10%) Intrinsic damping (10%) Damping = 75% Code wind speed, intrinsic damping (10%) Code wind speed, damping = 75% Wind speed from climate study, damping = 75% Factored overturning moment (GN m) 152 (cross-wind governs) 60 (cross-wind not considered) 50 74 45 37

Table 2. Variation of lateral acceleration with damping Effect Lateral acceleration: 10-year wind Lateral acceleration: 10-year wind Lateral acceleration: 10-year wind Method WTT WTT WTT Assumption Climate study windspeed, intrinsic damping = 10% Climate study windspeed, Damping = 75% Suggested limit for 10 year wind Lateral acceleration (milli-g) 256 94 15

Copyright 2007 John Wiley & Sons, Ltd.

Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

THE DAMPED OUTRIGGER CONCEPT FOR TALL BUILDINGS

515

Two dampers per wall were chosen rather than one, for the following reasons:

cheaper to supply; easier to install; greater redundancy.


Behaviour of damper The damper has a non-linear characteristic described below: F = C1v 2 At an upper threshold velocity, the behaviour changes to F = F1 + C2 v 0 1 This change in behaviour was chosen to prevent overload caused by extreme seismic demands and is implemented by the use of a pressure release valve. The threshold force F1 is the maximum force seen in the dampers during wind loading. At low levels of damper velocity the force mobilized is very low. This reduces the working pressure within the damper, increases the lifetime of the moving components and reduces susceptibility to fatigue. The damping system provides more damping when the wind loading approaches the maximum design levels. Damping achieved The additional damping achieved in each direction for 100-year wind varied between 52% and 112% of critical for the two buildings and two principal directions. Damper strokes and forces The peak stroke during the design wind storm was predicted to be approximately 50 mm. For the 2475-year return period earthquake this increased to 200 mm. The peak force was 22 MN per damper, with two dampers per outrigger. 7. ECONOMIC BENEFIT

In the case of this development, the construction cost saving (for construction in the Philippines) made by use of dampers, alongside performance-based seismic design, was approximately US $5 million. This cost saving is net of the cost of the dampers themselves. The quantity of concrete in the buildings was reduced by 30%, and the net oor area increased by about 2%. The reinforcement density in the towers was reduced from an average of 300 kg/m3 to approximately 200 kg/m3, this largely as a result of adopting performance-based seismic design. Benets of a similar order have been made on other projects using this technology. 8. CONCLUSIONS

This paper presents a new philosophy for the design of high-rise buildings where damped outriggers or variants thereof are used to introduce a high level of dependable supplemental damping for the primary purpose of reducing dynamic wind effects.
Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

516

R. J. SMITH AND M. R. WILLFORD

Signicant economic savings can be made by introducing damping in this way because the lateral stiffness of a high-rise building can be reduced, thereby reducing the element sizes and material content, without any adverse effect. In addition, the cost and space requirements for tuned mass dampers are eliminated. The application can be used to control response from both wind and seismic loading, but in both cases non-standard analytical techniques are required in the design process. The practicalities of designing and maintaining such a system are discussed and clear economic benets shown.
REFERENCES

Association of Structural Engineers of the Philippines. 2001. National Structural Code of the Philippines (NSCP). FEMA. 2000. Prestandard and Commentary for the Seismic Rehabilitation of Buildings, FEMA 356, Federal Emergency Management Agency: Washington DC. Higashino M, Okamoto S (editors), CIB Task Group 44. 2006. Response Control and Seismic Isolation of Buildings. Taylor and Francis, Oxford. McNamara RJ, Taylor DP. 2003. Fluid Viscous Dampers for High-Rise Buildings. Structural Design of Tall and Special Buildings. 12: 145154. National Research Council Canada. 2005. National Building Code of Canada (NBCC). Tamura, Y. 2006. Amplitude Dependency of Damping in Building and Estimation Techniques Keynote lecture, 12th Australasian Wind Engineering Conference, Queenstown, New Zealand, Jan 31Feb 2 2006.

Data sources for gure 1


Bouwkamp JG, Stephen RM. 1973. Ambient and forced vibration studies of a multistory pyramid-shaped building. Fifth World Conference on Earthquake Engineering, Rome. Brownjohn JMW. 2005. Lateral Loading and response for a tall building in the non-seismic doldrums. Journal of Engineering Structures 27: 18011812. Brownjohn JMW, Ang C-K. 1998. Full-Scale Dynamic Rsponse of High-Rise Building to Lateral Loading. Journal of Performance of Constructed Facilities, February: 3340. Brownjohn JMW, Pan T-C, Cheong H K. 1998. Dynamic response of Republic Plaza, Singapore. The Structural Engineer 76(11): 221226. Cheng CM, Kareem A. 1992. Acrosswind Response of Reinforced Concrete Chimneys. Journal of Wind Engineering and Industrial Aerodynamics 4144: 21412152. Cho KP, Tamura Y. 1999. Field Measurement of Damping in Industrial Chimneys and Towers. 1st Intl Conferences on Advances in Structural Engineering and Mechanics, Seoul. Ellis BR, Littler JT. 1988. Dynamic response of nine similar tower blocks. Journal of Wind Engineering and Industrial Aerodynamics 28: 339349. Fujii K, Tamura Y, Sato T, Wakahara T. 1980. Wind-induced vibration of tower and practical applications of tuned sloshing damper. Journal of Wind Engineering and Industrial Aerodynamics 33: 263272. Hill-Carroll P. 1985. The prediction of mean structural damping values and their coefcients of variation, ME.Sc Thesis, University of Western Ontario. Jeary AP. 1997. Damping in structures. Journal of Wind Engineering and Industrial Aerodynamics 72: 345355. Jeary AP. 1986. Damping in Tall Buildingsa Mechanism and a Predictor. Earthquake Engineering and Structural Dynamics 14: 733750. Jeary AP, Winney PE. 1972. Determination of structural damping of a large multi-ue chimney from the response to wind excitation. Proceedings of the Institution of Civil Engineering 53: 569577. Li QS, Fang JQ, Jeary AP, Wong CK, Liu DK. 2000. Evaluation of wind effects on a supertall building based on full-scale measurements. Journal of Earthquake Engineering and Structural Dynamics 29: 18451862. Li QS, Yang K, Zhang N, Wong CK, Jeary AP. 2002. Field measurement of amplitude-dependent damping in a 79-storey tall building and its effects on the structural dynamic responses. Structural design of tall buildings 11: 129153. Li QS, Fu JY, Xiao YQ, Li ZN, Ni ZH, Xie ZN, Gu M. 2006. Wind tunnel and full-scale study of wind effects on Chinas tallest building. Journal of Engineering Structures 28: 17451758.
Copyright 2007 John Wiley & Sons, Ltd. Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

THE DAMPED OUTRIGGER CONCEPT FOR TALL BUILDINGS

517

Li QS, Xiao YQ, Wong CK, Jeary AP. 2003. Field measurements of wind effects on the tallest building in Hong Kong. Structural design of tall and special buildings 12: 6782. Park W, Park K-S, Koh H-M, Ha D-H. 2006. Wind-induced response control and serviceability improvment of an air trafc control tower. Journal of Engineering Structures 28: 10601070. Satake N, Suda K, Arakawa T, Sasaki A, Tamura Y. 2003. Damping evaluation Using Full-Scale Data of Building in Japan. Journal of Structural Engineering, ASCE 129: 470477. Wood G, 2007. Damping measurements on high-rise buildings from University of Sydney. Private correspondence. Xu YL, Zhan S. 2001. Field measurement of Di Wang Tower during Typhoon York. Journal of Wind Engineering and Industrial Aerodynamics 89: 7393.

Copyright 2007 John Wiley & Sons, Ltd.

Struct. Design Tall Spec. Build. 16, 501517 (2007) DOI: 10.1002/tal

Potrebbero piacerti anche