Sei sulla pagina 1di 11

ARTICLE IN PRESS

Thin-Walled Structures 46 (2008) 720– 730

Contents lists available at ScienceDirect

Thin-Walled Structures
journal homepage: www.elsevier.com/locate/tws

‘Hybrid’ light steel panel and modular systems


R.M. Lawson a,, R.G. Ogden b
a
School of Engineering, University of Surrey, Guildford GU2 7XH, UK
b
Department of Architecture, Oxford Brookes University, Oxford OX3 0PB, UK

a r t i c l e in f o a b s t r a c t

Available online 18 April 2008 Modern methods of construction (MMC) are defined as those which are highly pre-fabricated and which
Keywords: achieve tangible benefits to the client in terms of speed of construction, higher quality and more
Light steel efficient and adaptable space use. There are many examples of MMC in light steel framing and modular
Modular construction, which are targeted on the residential and mixed-use building sectors. Modular units can
Demonstration building be designed with partially or fully open sides so that two or more modules can be placed side by side to
Testing create larger spaces. An alternative ‘hybrid’ approach is to combine 3D-modules for the highly serviced
and higher value parts, such as kitchens and bathrooms, and to use long span 2D-panels for the floors
and walls in the more open plan areas. The long span floor cassettes typically span up to 6 m between
separating walls or the sides of the modules. The floor cassettes occupy the same depth as the floor
and ceiling of the module and achieve a target depth of 450–500 mm. The paper reviews the design and
construction of a ‘hybrid’ demonstration building addresses the background development work and
testing of the modules and floor cassettes.
& 2008 Published by Elsevier Ltd.

1. Introduction long spans with service integration are well understood. The
medium-rise residential sector, such as apartments, hotels and
Pre-fabrication by off-site manufacture (OSM) leads to faster student residences uses similar steel and composite technologies,
construction, improved quality and reduced resources and waste. although at a more modest scale.
Although pre-fabrication is not in itself new, OSM describes a Modular construction is an example of a high level OSM, but
supply and construction process in which the major parts of a there are also opportunities for ‘hybrid’ planar and volumetric
building are mass-produced in factory conditions rather than on technologies, which optimise the value–cost balance in housing.
site. So-called ‘modern methods of construction’ (MMC) are ‘Open building’ systems are relatively advanced as they allow for
defined by their improvements in terms of the targets set by the interchange of components to create more flexible building forms
UK Government’s report [1]. than is achievable in fully modular construction. This is the area in
Steel construction is, by its nature, pre-fabricated to some which the greatest advances are possible, and a CIB Working Group is
degree, but the innovative use of this technology has arisen in currently exploring open building systems at an international level.
response to market demand for higher levels of pre-fabrication. In
the context of this paper, the uses of highly pre-fabricated
construction systems will be reviewed, showing how steel 1.1. Case examples
technology has developed over the last 5 years, and how basic
research information has been established to support these new Recent projects in UK have demonstrated the benefits of pre-
developments. fabricated construction technologies, such as the award-winning
The sector for which MMC is being promoted is in housing and Murray Grove project in Hackney, London, completed in 1999,
residential buildings, which also includes single person accom- which used modular construction based on the Yorkon system.
modation and affordable housing, particularly in inner cities. Steel More recently, the Lillie Road project in Fulham, west London,
construction has established a ‘track record’ in the commercial completed in 2003, used light steel framing, modular bathrooms
building sector, where the benefits of speed of construction and and a slim floor primary frame at first floor to optimise both the
construction process and provision of space for this mixed-use
building. In both projects, the client was The Peabody Trust, which
 Corresponding author. Tel.: +44 1483 686617. took a strong interest in realising the value-benefits of these
E-mail addresses: m.Lawson@surrey.ac.uk (R.M. Lawson), relatively new technologies. These projects are illustrated in
rgogden@brookes.ac.uk (R.G. Ogden). Figs. 1 and 2.

0263-8231/$ - see front matter & 2008 Published by Elsevier Ltd.


doi:10.1016/j.tws.2008.01.042
ARTICLE IN PRESS

R.M. Lawson, R.G. Ogden / Thin-Walled Structures 46 (2008) 720–730 721

Fig. 3. Royal Northern College of Music, Manchester, consisting of 900 modules


and braced steel cores.

Fig. 1. Installation of modular units at Murray Grove, Hackney, London.

Fig. 4. Mixed commercial–residential development at Wilmslow Road, Manche-


ster, using 1400 modules on a steel-composite podium structure.
Fig. 2. Completed mixed panel and modular project at Lillie Road, Fulham, London.

The world’s largest modular buildings are located in Manche- order to be fully economic. OSM requires capital investment in the
ster and use a similar technology based on the Ayrframe system, infrastructure of factory production, design development, product
an innovative form of ‘stressed skin’ construction. The Royal testing and certification, and overheads of a fixed facility and
Northern College of Music student residence consists of 900 factory space. Cellular-type buildings, such as hotels and student
modules in a 6–9-storey configuration (see Fig. 3), and a mixed residences have multiple similar units, and are the types of
communal-retail development consists of 1400 modules sup- projects where OSM has proved to be successful. The break-
ported on a 2-storey podium in composite construction (see through of OSM into the wider residential sector is still in its
Fig. 4). infancy.
Unite Modular Solutions, a major ‘design build finance operate’ Modern highly automated factories for modular production
provider in the student residence and key worker sector, has cost of the order of h15 million to set up. Although much less than
completed many projects using fully modular construction, and the h1000 million required to set up a new automotive production
has commissioned a new factory to produce bedroom modules at line, these costs are distributed over a yearly output of 1000–2000
a target rate of up to 20 per day. Other initiatives are underway in units in a changeable building market, in comparison to a typical
UK, involving a variety of modular and panel systems, notably by annual production of 50,000 of a successful car model over a
Advance Housing, involving collaboration of Terrapin and a major 7-year cycle. Balanced against these fixed capital costs are savings
house builder, and by Kingspan Off-site. due to more efficient production technologies, reduced site
construction costs, higher quality levels and time-related savings
1.2. Economics and production of offsite manufacture due to speed of construction. Although it is recognised that time
savings of 30–50% in total construction time can be realised by
The underlying economics of off-site manufacturing (OSM), modern OSM, the economic value of this early completion
and modular construction in particular, is quite complex and depends on the business operation or early sales revenue. This
requires a significant production rate of repeatable components in can be quantified for a hotel chain or a time-constrained
ARTICLE IN PRESS

722 R.M. Lawson, R.G. Ogden / Thin-Walled Structures 46 (2008) 720–730

Fig. 5. Assembly of light steel panels in production.

Fig. 6. Light steel framing for housing.

operation, such as a university, but is less apparent for a house


builder in a speculative market.
Essentially, the additional costs of a permanent factory have to
be balanced against savings in inefficient and wasteful site
operations. Most OSM projects involve a proportion of site work
(20–40% being typical), which are reflected in the broad costs in
this figure. Although OSM leads to efficiencies in materials use
and reduced wastage, many pre-finished components are bought
in, which increases their cost. Small OSM projects may not result
in significant economies, unless the same form of construction is
repeated in a number of similar projects. However, large OSM
projects can lead to cost savings of 10–20% in addition to time
savings by reducing site infrastructure costs and increasing
productivity and reliability.
The rationale behind the expansion of OSM depends on
investment in numerically controlled machinery and integrated
CAD/CAM software. In Europe, parallel technology was first
developed in the timber frame industry, whereas in Japan,
companies such as Sekisui and Toyota Homes are advanced in Fig. 7. Continuously supported module in light steel framing (by Terrapin).
implementation of steel-based technologies in modular construc-
tion. Light steel sections may now be produced by small-scale paths through the walls by supporting the floors, for example, on a
roll-forming machines, and panels are assembled accurately on Z trimmer attached to the top of a wall panel. This technology is
tables and boards are fixed rapidly, for example, using ballistic described by Gorgolewski et al. [2].
nailing. A typical factory assembly process is shown in Fig. 5. Up to
30 stages are required in a continuous modular production
facility. Completed modules are sent directly to site for ‘just in 2.2. Modular construction
time’ delivery.
Volumetric or modular construction systems are manufactured
from 2D wall panels and floor cassettes in light steel framing, but
2. Generic forms of light steel and modular construction are assembled into load-bearing ‘boxes’ which are fitted out and
transported to the construction site. The primary limitations are
Historically, steel has been used in housing for 70 years, and those of production and transport as factory manufacture requires
there are many good examples of its use worldwide. The modern multiple similar units, and transport necessitates a unit width of
forms of steel and mixed construction systems that are widely less than 4.1 m. This technology is described by Lawson et al. [3].
used in the housing and residential sector are described in simple Two generic forms of modular construction exist:
terms as follows:
 Continuously supported or 4-sided modules where vertical
2.1. Light steel framing: elemental and panel systems loads are transmitted through the walls (Fig. 7).
 Open-sided or point-supported modules where vertical loads
Light steel framing consists of galvanised steel C-sections of are transmitted through corner and intermediate posts (see
typically 65–200 mm depth and in steel thicknesses of Fig. 8).
1.2–2.4 mm. Walls are generally pre-fabricated as 2D-panels, as
in Fig. 6, whereas floors can be installed in elemental form as Point-supported systems require deeper edge beams than con-
joists or in 2D-cassette form. For 2-storey buildings, platform tinuously supported modules. In both systems, resistance to
construction may be used (i.e. floors sit directly on walls) but for horizontal loads can be provided by bracing or diaphragm action
medium-rise design, it is necessary to achieve continuity in load in the walls, but for buildings more than 6-storey high, a separate
ARTICLE IN PRESS

R.M. Lawson, R.G. Ogden / Thin-Walled Structures 46 (2008) 720–730 723

Fig. 8. Corner-supported module (by Kingspan). Fig. 9. Demonstration building using mixed panel and modular construction.

bracing system is required, which is often provided around the


access core. Forces are transferred by the module–module
connections in the form of plates and bolts, assisted by horizontal
bracing in the corridors.

2.3. ‘Hybrid’ modular and panel systems

‘Hybrid’ or mixed modular and panel systems optimise the use


of the 3D and 2D components in terms of space provision and
manufacturing costs. Modular units are used for the higher value
Fig. 10. Hybrid panel and modular construction (by courtesy of Corus).
of more highly serviced areas, such as bathrooms, wall panels and
floor cassettes for the more flexible open space. Two generic forms
of ‘hybrid’ construction may be considered:

 Load-bearing modules with floors supported by the modules.


 Non-load bearing modules (or pods) supported by floors.

The first system was used in a demonstration building for Corus,


shown in Figs. 9 and 10, in which the central service core and
stairs were manufactured as modules and the open plan space
was provided by pre-fabricated panels and floor cassettes spanned
up to 5.7 m. In this way, the internal space could be partitioned to
suit the user’s requirements. The building form is presented
by Lawson et al. [4]. The construction of the Lillie Road project
(in Fig. 2) comprises X-braced wall panels, floor cassettes and
stacked bathroom modules, as shown in Fig. 11.

2.4. ‘Hybrid’ modular, panel and primary steel frame

Fig. 11. Load-bearing braced walls and stacked bathroom modules in the project
Modular construction has so far only been used for medium-
shown in Fig. 2.
rise cellular buildings. Greater flexibility in building height and
internal planning can be achieved by the mixed use with a
primary steel structure. Various generic forms of construction
may be employed by creating: A podium structure is often used where retail outlets or
communal space are provided at ground floor and car parking in
 A ‘podium’ structure of typically one or two storeys height in the basement, as in the project shown in Fig. 4. Composite
which the column spacings are located at multiples (two or construction may be used in which the podium level is designed
three times) the module width. to support the load from the modules above (typically 6 storeys).
 A skeletal structure, which provides the open plan areas and A skeletal structure may be designed in the form of slim floor
the stacked modules provide the highly serviced areas or cores. beams using HE or RHS sections in which the modular, and
 A skeletal structure, in which non-load bearing modules and floor cassettes are supported on the extended bottom flange so
wall panels are supported on the floor. that the beams occupy the same depth as the floor. A pair of
ARTICLE IN PRESS

724 R.M. Lawson, R.G. Ogden / Thin-Walled Structures 46 (2008) 720–730

7.2 m

3.6 m Floor span

Internal angle
Open side
0.3 m Corne
Site infill
angle

Open side
≤ 3.0 m Site infill

Floor span Floor span


Fig. 12. Recessed modules supported by primary steel frame.
Internal angle

modules would be located within the column grid, and the


corners of the modules are recessed in order that they fit around 3.6 m Internal wall
SHS or narrow columns in order to minimise wall widths, as
shown in Fig. 12.

2.5. Open-building systems ≤ 3.0 m 0.3 m ≤ 3.9 m

‘Open-building’ technology is a general term used to describe Fig. 13. Creation of flexible space using modules.
systems, which provide flexibility in space planning and in
interchange of components. Many of the hybrid systems described
above achieve some of the principles of ‘open technology’, but to
be more widely applicable and to achieve economy in manufac-
ture, geometrical standards and common interface standards are
required for the cladding, services, lift and stairs and other key
components. Geometric standards that may be used for concept
design which are based broadly on the following dimensions:

 Wall width of 300 mm for internal separating walls and


external walls.
 Floor depth of 450 mm for the combined floor and ceiling
depth in modular and ‘hybrid’ construction systems.
 Floor depth of 600 mm when a supporting primary steel
structure is used.
 Internal planning dimensions based on 600 mm on plan
(therefore 3 or 3.6 m are preferred internal modular widths.
 Floor-ceiling heights based on 2.4 m for residential buildings
and 2.7 m for commercial, health or educational buildings.

Modular construction achieves the benefits of OSM, but it requires


a new discipline in construction technology based on building
‘blocks’ rather than skeletal or planar components with which
designers are familiar. An optimised modular system must allow
for greater flexibility in internal planning, but must retain the
primary benefits of speed of installation and improved quality.
The inter-relationship between modules and efficient provision of
space can be improved by strategically placed internal posts,
which allow for both open-sided design and for re-orientation of
modules. A typical plan of such a group of modules is illustrated in
Fig. 13. Openings of up to 3 m width can be created, and a cluster
of posts form a column which can support loads of up to 8 storeys.
OpenHouse AB is a Swedish system in which recessed modules
are supported on a grid of 3.9 m by SHS columns, as illustrated in Fig. 14. Installation of modules with recessed corners around SHS columns
Fig. 13. OpenHouse is described by Lessing [5] (Fig. 14). (courtesy of OpenHouse AB).
ARTICLE IN PRESS

R.M. Lawson, R.G. Ogden / Thin-Walled Structures 46 (2008) 720–730 725

2.6. ‘Hybrid’ demonstration building sheet. The building was constructed in only 4 weeks. An
interesting technology that was explored was the use of long
A demonstration building using mixed modular and panel span floor cassettes using timber-steel composites, as shown in
construction was designed and built in 2003 as a way of investi- Fig. 16. These floor joists were subject to extensive testing, and
gating the application of this new technology. In this approach, this work has been continued with further tests at the University
the load-bearing elements are the modules and separating of Surrey. The flanges comprise 56  1.6 C sections and the webs
walls, which minimises the foundations, and the modules are are in the form of 12 mm plywood that is ballistically nailed to the
braced for stability up to 5 storeys high. The space can be flanges. The flanges resist tension and compression and the nailed
partitioned to suit the desired room sizes. The demonstration web resists shear to a resistance of approximately 10 kN/m, which
building is illustrated in plan form in Fig. 15. It is 14 m wide and is suitable for residential applications.
10 m long and is part of an ‘urban terrace’. Although the All separating floors had 18 mm chipboard over 40 mm Rock-
demonstration building is only 2-storey high, it is extendable to floor with 19 mm plasterboard or 18 mm chipboard attached to the
taller buildings. joists, with the layers on 12.5 mm fire-resistant plasterboard with
A range of acoustic, thermal insulation and floor vibration tests 100 mm mineral wool insulation placed between the joists.
was performed on the demonstration building on its completion. Thermal analyses were carried out for the two wall types in the
The completed demonstration building is illustrated in Fig. 9. demonstration building: an insulated render, and a brick-tile
Its cladding is in the form of an insulated render or cement system (Corium). The computed thermal transmissions (U-values)
particle board, or a brick slip system on a metallic ribbed backing for various thicknesses of insulation are presented in Table 1.

D 3 (Floor)
D 137
1075
1075 237 1075 300 100
D 3
2469 2876

2717 F F
3861
A A
2 2
2 2 Floor
(Floor)
100 4983

126
100

10293 External
10019 Internal

600 100

2403 152

E E C C
2049

4234
152

300 4936

2429 100

B B 2695 1 1 (Floor)

137 137 3 (Floor)

3
7051

Fig. 15. Plan form of demonstration building.


ARTICLE IN PRESS

726 R.M. Lawson, R.G. Ogden / Thin-Walled Structures 46 (2008) 720–730

ceiling, the long span floor joists can be approximately 320 mm


deep and have span capabilities up to 5.7 m.
The modular floor with its perimeter C section supports the
long span floor, but the bending resistance of the floor cassette is
provided by a 320 mm deep  3 mm thick C section included its
manufacture. The floor cassette may be up to 3 m wide depending
on transportation and installation requirements.
The long span floor cassette is supported by a continuous steel
angle or Z section attached to the lower module or separating
wall, as illustrated in Fig. 19. In this way, the walls of the modules
provide a direct load path.
This concept may be adapted to other building forms, where
the kitchen/bathroom modules are arranged transversely within
the plan form, and the open space is provided next to the fac- ades
for improved daylight. Service risers may be combined in the
adjacent modules.

3. Background testing

As a background to this demonstration project, two series of


tests on light steel wall panels, acting in shear, were carried out to
investigate composite actions with:

 Different sheathing boards


 Brickwork cladding.

3.1. Shear tests on wall panels

Shear (racking) tests were carried out at the University of


Surrey and Corus RD&T to investigate the performance of different
Fig. 16. Timber–steel composite joists used in demonstration building. sheathing materials on a plain light steel wall panel and a panel
with a 1.2 m square window. The chosen sheathing materials for
the tests were:
Table 1
Thermal insulation for different wall constructions
 Plasterboard (internally)—12.5 mm wall board.
Description 2
U-value (W/m K)  Plywood (externally) and plasterboard (internally)—12 mm
plywood and 12.5 mm wall board.
Render: 80 mm EPS, Rockwool in frame cavity 0.236  Cement particleboard (externally) and plasterboard (intern-
Render: 100 mm EPS, Rockwool in frame cavity 0.206 ally)—10 mm cement particle board.
Corium: 50 mm EPS, Rockwool in frame cavity 0.300
Corium: 80 mm EPS, Rockwool in frame cavity 0.235
 Steel sheeting (externally) and plasterboard (internally)—the
Corium: 100 mm EPS, Rockwool in frame cavity 0.205 steel sheeting was 19 mm deep, 0.55 mm thick with an 84 mm
pitch and spanned horizontally between the wall studs.

The wall shear tests were carried out 2.4 m square wall frames
Mineral wool (Rockwool) placed between the wall studs reduces with a riveted light steel frame using C sections of 75 mm depth
the U-values considerably. and 1.6 mm thickness in S350 steel. These light steel frames are
The thermal performance of the rendered cladding system typical of general applications in housing. Vertical wall studs were
with mineral wool between the wall studs is shown in Fig. 17. located at 600 mm centres and plain C horizontal rails were fixed
The isothermal lines show the local ‘hot spots’ on the steel studs. across the top and bottom of the panels. A minimum of four self-
The maximum temperature difference is 2 1C across the wall piercing rivets was used at the end of each stud, two per flange
(for a 20 1C overall temperature between inside and outside). The connection.
use of brick tiling shows a similar performance, increasing the The plasterboard was fixed using standard screws at 300 mm
amount of external insulation reduces the local ‘hot spots’ to less centres. For the initial series of serviceability tests, sheathing
than 1 1C. boards were also fixed with screws at 300 mm centres, but for the
The ‘hybrid’ modular-panel concept uses highly serviced later tests, additional fixings were introduced to reduce the
modules for the kitchen, bathrooms, stairs and lift as a ‘core’ of spacing to 150 mm for the ultimate load tests. All holding down
the building and long spanning floor cassettes supported by the arrangements had standard bolted brackets at the corners and
modules and separating walls. In this way, flexible space is created base of the panel.
which can be fitted out to suit the user’s requirements. This The panels were all tested subject to in-plane loads and the
‘hybrid’ concept is illustrated in Fig. 18. test procedure outlined in BS 5268 Section 6.1 was used
The combined depth of the floor and ceiling of the module is throughout the tests. The stiffness test applies an increasing
approximately 400 mm, and so the floor cassette can be of the lateral load to the head of the frame until a net horizontal
same depth. Allowing for acoustic build-up and the plasterboard deflection of 4.8 mm is measured at the top corner of the
ARTICLE IN PRESS

R.M. Lawson, R.G. Ogden / Thin-Walled Structures 46 (2008) 720–730 727

Fig. 17. Thermal profile—Render, 60 mm EPS, 100 mm Rockwool between C sections.

2.4 2.4

Balcony 1.2 Bynocla

Kitchen/ Kitchen/
bathroom bathroom 3.8
module module

Separating wall Flexible 10.8 m


Stair space
4.2
module (50 m2 approx)

Lift/ 2.8
lobby
module

0.6
4.9 3.2
6.5 m Frontage

Fig. 18. Plan form of ‘hybrid’ building for adjacent apartments.

frame, which corresponds to a serviceability deflection limit of stiffening effects of vertical load are relatively small, unlike in
height/500. timber framing where tensile action at the base of the wall studs
All tests were carried out without applied vertical loads for dominates. The horizontal load was then increased until failure,
both stiffness and strength tests. In light steel framing, the and generally, the failure load was at least twice the serviceability
ARTICLE IN PRESS

728 R.M. Lawson, R.G. Ogden / Thin-Walled Structures 46 (2008) 720–730

load. Therefore, it is the stiffness criterion, which controls the The influence of vertical load is apparent from Table 3, which
design of the panel. included tests on plain panels, and panels with window openings
Table 2 summarises the results of the tests for the light steel using plasterboard or Fermacell board. A vertical load of 5 kN/m
wall panels with and without windows. The cement particleboard corresponds to factored loading from a single floor, and a modest
provides the greatest increase in shear resistance of the panel. load of half this value is sufficient to increase the shear stiffness
When compared to the results for the plasterboard-clad panel, the of the wall panel by 10–15%. Cross-flat bracing is shown to be the
increase in design load was 95% for plywood, 57% for the steel most efficient, but this is not a practical solution for panels with
sheeting and 143% for cement particle board. When the fixing windows.
spacings were reduced from 300 to 150 mm, the failure load Design (unfactored) shear loads are taken as either the average
increased by an additional 11%, 30% and 21%, respectively, for the value determined from the serviceability tests, or the minimum
three board materials. The increase in stiffness of the steel test load divided by a factor of safety of 1.6. It should be noted that
sheeting is due to the reduced deformation of the sheet profile for BS 5268 Section 6.1 requires a reduction factor of 0.8 for single
closer spaced fixings. tests, but allows a multiplication factor of 1.25 to represent a

20
30
20 20

410 320

12
30

310 x 100 x 3 C

150 x 100 x 10 L
150

330 ext.
150

30

Fig. 19. Attachment of floor cassette to modules in ‘hybrid’ construction: (a) cross-section through floor and module and (b) detail of edge beam to floor cassette.

Table 2
Tests on wall panels with various forms of diaphragm action

Configuration of 2.4 m square wall panel Service load (kN) based on stiffness Failure load (kN) Design load (kN/m) zero vertical load

Spacing of fixings at perimeter of boards (mm) 300 150 150 300 150

No openings
Plasterboard (Pb) 3.7 – – 1.5 –
Plywood and Pb 7.2 8.0 26.8 3.0 3.3
Cement particleboard and Pb 9.0 11.0 34.6 3.7 4.6
Steel sheeting (19 mm) and Pb 5.8 7.5 23.8 2.4 3.1

With window opening


Plasterboard (Pb) 3.0 – 13.5 1.2 –
Plywood and Pb 6.0 8.6 2.5 3.6
Steel sheeting (19 mm) and Pb 4.0 4.4 18.9 1.6 1.8

Note: design load is the unfactored wind load resisted in shear.


ARTICLE IN PRESS

R.M. Lawson, R.G. Ogden / Thin-Walled Structures 46 (2008) 720–730 729

Table 3
Shear load (in kN) for wall panels showing the influence of vertical load

Configuration of 2.4 m square wall panel Horizontal service load based on stiffness for: vertical Failure load (kN) Design load (kN/m)
loading on wall (kN/m)

0 2.5 5 Zero vertical load Zero vertical load

Plain panel: single integral bracing 4.9/5.0 5.5/5.6 5.5/5.7 13.3 2.0
Plain panel: cross flat bracing 8.7/8.8 9.6/9.7 10.2/10.3 25.4 3.6
Window panel: double integral bracing 3.8/4.0 4.4/4.5 4.3/4.4 12.2 1.6
Window panel: no bracing. Fermacell board 5.9/6.0 6.5/6.7 7.0/7.2 13.7 2.4

stiffness enhancement corresponding to a shear deflection of Table 4


0.003  height for timber framing. The multiple of these two Shear load (in kN) for 2.4 m square brick-clad panels at a deflection limit of
height/500
factors is clearly unity in this case.
Frame type Without With brickwork and plasterboard
3.2. Wall panel tests attached to brickwork brickwork
Frame with 35 mm thick 70 mm thick
plasterboard insulation insulation
Tests were conducted at and reported by Ceram Research [6]
on 2.4 m square light steel wall panels attached to brickwork to 1. Plain panel—no 3.7 10.8 (7.3) 8.1
evaluate their enhanced shear stiffness. Four different types of bracing
2. Plain panel—single 5.1 9.4 (8.9) 7.9
light steel frame were used in the tests: bracing
3. Plain panel—double 5.5 11.3 10.6
1. Plain panel—no in-plane bracing bracing
4. Window 2.7 11.4 7.4
2. Plain panel—single bracing line integrated between the studs
panel—double
3. Plain panel—double bracing lines bracing
4. Window panel—double bracing lines.
Notes: value in brackets indicates a test on a brickwork panel without plasterboard.

The panels had 12.5 mm thick plasterboard fixed by screws at


300 mm centres. On the opposite face, the brickwork was
connected to the frames by flat wall ties, which had a minimum 4. FE model of wind forces on K-braced frame
vertical spacing of 375 mm, reducing to 225 mm next to
openings. The wall ties fitted into vertical steel channels, In principle, K-braced panels act as vertical cantilevers and
which were attached to each stud (located at 600 mm centres). react against the foundations by developing tension–compression
The wall tie density was approximately 4.4/m2. As this type of forces. However, the braced fac- ade can act beneficially as a
panel is used in a ‘warm frame’, the channels were fixed whole by local bending in the wall studs and in the lintels,
by long screws through the external insulation, which affects so that forces in the bracing members and wall studs are
the stiffness of the wall tie system. The wall ties were flat less than given by the above simple analysis on a single K-braced
stainless steel ties of 42 mm width and 1.6 mm thickness. When panel.
bonded into the brickwork, the ties are relatively stiff in the A finite element (FE) model of a typical fac- ade with four
horizontal direction, but are flexible in the vertical direction. K-braced panels was carried out, subject to horizontal forces of
The distribution of the ties is important because of their effect 10 kN at each floor level (or a base shear of normally 5 kN per
in transferring horizontal forces between the panel and the K-braced panel). The applied over-turning moment on the wall is
brickwork. therefore 78 kN/m for a storey height of 2.6 m. Dividing by four
The panels were constructed within a rigid steel frame, K-braced panels and a distance of 0.6 m between the wall studs
which held the brickwork in position during the test, while the leads to a tension-compression force of 32.5 kN less than the self-
light steel frames were free to displace when subject to weight of the fac- ade (taken as a nominal 2 kN/m).
an in-plane point load. The wall was constructed using solid clay The FE model leads to the member forces illustrated in Fig. 20.
bricks and mortar to designation (iii) to BS 5268. The test The maximum forces experienced in the lower part of the wall
programme covered tests with single (35 mm) and double stud have a maximum value of 21.8 kN, which is equivalent to 67%
(70 mm) layers of closed-cell polyurethane insulation. The cavity of the simplified cantilever method based on the over-turning
space between the insulation and the brickwork was 50 mm in moment applied to the four K-braced panels. This suggests
all cases. that structural action of a group of K-braced panels reduces the
The results of these wall panel tests are presented in Table 4. In compression forces by approximately 33%. Furthermore, the
all cases, serviceability (stiffness) is the controlling design tension forces in the wall studs are considerably less and
condition (see previous explanation). The attachment of the maximum value is 14.7 kN (or only 45% of the simplified
brickwork with a single layer of 35 mm thick insulation doubles cantilever method). It follows that the simplified over-turning
the shear resistance compared to a frame with only plasterboard. model is very conservative for an integrally K-braced fac-ade.
The test results are relatively independent of the panel type,
indicating that the stiffening effect of the brickwork is dominant,
but the stiffening effect of the brickwork was 10–30% less 5. Conclusions
when two layers of insulation were used. Nevertheless, the
shear load that can be resisted by the brick-clad panel was This paper reviews modern methods of light steel construction
increased by a factor of over 2.5 relative to a braced wall panel that are used in the residential sector, and identifies mixed
with a window. ‘hybrid’ forms of planar and volumetric construction that are
ARTICLE IN PRESS

730 R.M. Lawson, R.G. Ogden / Thin-Walled Structures 46 (2008) 720–730

Fig. 20. Finite element analysis of forces in bracing and wall studs for 10 kN shear forces applied at each floor in a 2-storey house.

economic in the medium-rise sector. Tests on wall panels in shear References


are summarised without and with brickwork cladding. The
structural behaviour of light frames demonstrates considerable [1] Re-thinking construction. The report of the Construction Task Force (The Egan
reserve in overall stability. Report) published by the Office of the Deputy Prime Minister, UK; 1998.
[2] Gorgolewski MT, Grubb PJ, Lawson RM. Building design using cold formed steel
sections: light steel framing in residential buildings. The Steel Construction
Institute, P301, Ascot, UK; 2001.
[3] Lawson RM, Grubb PJ, Prewer J, Trebilcock PJ. Modular construction using light
Acknowledgements steel framing: an architect’s guide. The Steel Construction Institute, P-271,
Ascot, UK; 1999.
[4] Lawson RM, Ogden RG, et al. Steel in residential buildings for sustainable and
The research was carried out by the Steel Construction
adaptable construction. Final report to ECSC project 7215-PP-058, European
Institute and Corus under various European funded research Commission; 2004.
projects. Photographic material and additional information were [5] Lessing J. Industrial production of apartments with steel frames. A study of the
provided by Feilden Clegg Bradley, Michael Barclay Partnership, OpenHouse system. The Swedish Institute of Steel Construction Report 229-4,
Stockholm; 2004.
Kingspan Off-site, Corus Living Solutions, Terrapin, Yorkon and the [6] Testing of brickwork panels attached to light steel frames. Ceram research
Swedish Institute of Steel Construction. report SW181.98, Stoke on trent, UK; 1998.

Potrebbero piacerti anche