Structural Design of High-Rise Buildings Using Steel-Framed Modules - A Case Study in Hong Kong

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Received: 20 August 2019 Revised: 18 June 2020 Accepted: 28 June 2020

DOI: 10.1002/tal.1788

RESEARCH ARTICLE

Structural design of high-rise buildings using steel-framed


modules: A case study in Hong Kong

Sidi Shan | Wei Pan

Department of Civil Engineering, The


University of Hong Kong, Pokfulam, Hong Summary
Kong Steel-framed modular buildings afford certain advantages, such as rapid and high-
Correspondence quality construction. However, although steel-framed modules have been adopted in
Wei Pan, Department of Civil Engineering, several countries, most of them are limited to low-to-medium-rise structures; modu-
The University of Hong Kong, Pokfulam,
Hong Kong. lar high-rise buildings are rare. This study proposes a feasible structural design solu-
Email: wpan@hku.hk tion for high-rise buildings using a steel-framed modular system. A 31-story student
Funding information hostel building in Hong Kong is redesigned as a steel-framed modular building and
Research Impact Fund of the Hong Kong used as a case study. The finite element models of the building are formulated, and
Research Grants Council, Grant/Award
Number: R7027-18; Development Bureau of the structural behaviors under wind and earthquake load scenarios are compared.
the Government of the Hong Kong Special Moreover, the structural design process used for the 31-story building is applied to
Administrative Region, Grant/Award Number:
200008191 design a hypothetical 40-story modular building to further examine the proposed
design solution. The numerical analysis results indicate that the roof lateral displace-
ments and interstory drift ratios of the redesigned modular building are within the
allowable limits of design codes; moreover, the modular connections behave elasti-
cally under the most adverse loading scenarios. Accordingly, the proposed solution
can be used to design steel-framed modular buildings of up to 40 stories, while com-
plying with relevant wind and seismic codes.

KEYWORDS

finite element analysis, high-rise building, modular building, modular integrated construction,
steel-framed module, structural design

1 | I N T RO D UC TI O N

In many developed economies, the construction industry is confronted with severe problems, including aging workforce, labor shortage, and
increasing costs[1]; as a result, it is difficult to achieve a sustainable future for the industry and community. Modular construction
(i.e., construction using modules manufactured off-site) is a new approach, which may offer a solution: It affords advantages, such as shortened
on-site construction period, improved site health and safety, reduced reliance on on-site labor, minimal construction waste, and enhanced con-
struction quality.[2,3] If achieved, these benefits can aid in overcoming the problems confronting the construction industry. In Hong Kong, the use
of the modular approach in construction, particularly for high-rise buildings, has been promoted by the government under the “modular integrated
construction” (MiC) policy initiative.[4]
Steel-framed modules adopt structural steel frames as the structural system that can be facilely manufactured off-site and installed on-site to
construct modular buildings. To determine the global behaviors of modular buildings under various wind and earthquake loading scenarios, numer-
ous investigations have been performed over the past decade. Lacey et al.[5] presented an overview of the structural behaviors of low-to-
medium-rise modular buildings. Annan et al.[6] experimentally investigated the performance of a low-rise modular steel frame under hysteretic
loading and reported that the modular steel frame exhibited a similar ductility and energy dissipation capacity as those of regular steel frames.

Struct Design Tall Spec Build. 2020;29:e1788. wileyonlinelibrary.com/journal/tal © 2020 John Wiley & Sons, Ltd. 1 of 20
https://doi.org/10.1002/tal.1788
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2 of 20 SHAN AND PAN

Hong et al.[7] experimentally studied the cyclic mechanisms of modular steel frames with double-skin steel panels and found that these panels
enhanced the initial lateral stiffness of the frames. Annan et al.[8] and Fathieh and Mercan[9] examined the behaviors of four-story modular steel
frames and reported that the low-rise modular building exhibited a predominantly first-mode response under seismic scenarios. Gunawardena
et al.[10] investigated the lateral behaviors of medium-rise modular buildings subjected to seismic excitations. Srisangeerthanan et al.[11] confirmed
that a higher diaphragm flexibility significantly increased the interstory drifts of modular buildings during an earthquake. However, most of the
aforementioned studies focused on the overall performance of low-to-medium-rise modular buildings; only a few focused on high-rise modular
buildings although the behaviors of the latter significantly differ from those of the former under seismic-load and wind-load scenarios.
In recent years, numerous experimental,[12] numerical,[13,14] and analytical[15] studies on high-rise buildings have been performed; however,
only a few have considered modular high-rise buildings. Liu et al.[16] proposed a structural design method for a prefabricated high-rise steel struc-
ture; however, the structure was primarily composed of a prefabricated modular floor system and steel columns, rather than steel-framed mod-
ules. Because the investigations of the structural solutions and applications of high-rise structures with steel-framed modules are rare, it is
extremely important to ensure their structural stability and accurately ascertain their capacity.
The structural performance of high-rise buildings with steel modules is closely related to the mechanical behaviors of structural connections.
Many studies have investigated the performance of modular buildings using different connections. Lawson et al.[17] recommended a bolted-end
connection between two fully open-sided modules. Choi et al.[18] demonstrated that the lateral stiffness and strength of modular buildings may
be overestimated if the module-to-module connection is assumed to be rigid in the analysis. Doh et al.[19] investigated the behavior of a steel
bracket connection under shear and simply supported loading and reported that a ductile failure occurs in the connection when the tensile
strength of bolts is exceeded. Gunawadena[20] proposed a module-to-module connection and studied its behaviors under lateral loadings. Chen
et al.[21] presented a modified module-to-module connection and investigated its loading mechanisms; they demonstrated that stiffeners could
significantly enhance the stiffness and load-carrying capacity.[21] However, unless extensive experimental and parametric studies are performed,
modular connections are subject to various uncertainties. To avoid these uncertainties under various loading scenarios, it is necessary to imple-
ment capacity design to ensure that the modular connections in a modular building always behave elastically.
The objective of this study is to develop a structural solution for high-rise buildings using steel-framed modules (modular high-rise building)
that can ensure structural stability and safety. A real 31-story reinforced-concrete (RC) student hostel building is selected as the reference build-
ing and redesigned using steel-framed modules. A finite element (FE) simulation is performed to analyze the global behaviors of the reference and
redesigned modular buildings. Their lateral displacements, story drifts, and story shears under wind and seismic loads are compared. A scenario-
based solution is identified and proposed to ensure the structural stability and safety of the redesigned modular building. Moreover, to ensure
that the connections in the modular building behave elastically under the worst loading scenarios, the stress in these connections is determined.
Finally, the solution proposed in this study for the 31-story building is further examined by applying it to a hypothetical 40-story building.

2 | C A S E S T UD Y

2.1 | Reference building

In this study, a real 31-story (with a tower and podium) RC student hostel building is selected as the reference building (Figure 1). The building has
five podium stories and 26 residential tower stories. The building ground type (block A in Figure 1) is classified as Ground Type A.[22] The building
was constructed in 2014 via a conventional cast in situ RC construction scheme. The tower base measures approximately 25.7 × 15.2 m, and the
structure stands at a height of 82 m above the transfer level. The five podium stories below the transfer structure (podium level) are not consid-
ered in this study. The lateral force-resisting system mainly consists of a structural core wall system. The typical story height is 3.15 m, and the
slab thickness in each floor is 150 mm. The coupling beams range from 160 × 750 mm to 350 × 900 mm (width × depth), and the shear walls have
thicknesses ranging from 170 to 400 mm. The concrete grade is C45 (i.e., measured characteristic cube strength is 45 MPa), and the yield strength
of the reinforcement is 500 MPa. The distribution of gravity loads is illustrated in Figure 1b. The design wind load is calculated according to the
Hong Kong design code.[23] By dividing the total gravity load by the building volume, the building density is calculated as approximately
5.0 kN/m3.

2.2 | System scheme of high-rise building using steel-framed modules

The structural solution principle for the modular building is to satisfy all local building structural performance requirements with minimal changes
in the original architectural design. For the redesigned modular building, the common areas (e.g., elevator lobby, lavatories, and corridor area) are
assumed to be constructed via cast in situ concrete core wall system. The core wall system consists of cast in situ RC shear walls, coupling beams,
floor slabs, and so on. The residential areas are assumed to be constructed using steel-framed modules, each of which is suitable for a student
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SHAN AND PAN 3 of 20

FIGURE 1 The reference building. (a) Aerial view; (b) a typical floor plan

room. The modules are stacked and horizontally attached to the two longer sides of the cast in situ concrete core wall. The ceiling of each module
is specially designed to have a sufficient capacity to sustain horizontal loads (e.g., wind loads) and transfer the horizontal load to the core wall sys-
tem. The border plan of the core wall and modular system for a typical story is shown in Figure 2.
The columns and beams of each module are fabricated using rectangular hollow section (RHS) steel tubes. The side faces of each module are
covered with corrugated steel sheets with necessary openings to create door and window openings. The walls, ceiling, and floor are also
reinforced by steel bracings with specifically designed modular connections; the conceptual framing design for the steel modules is presented in
Figure 3. The module shown in the figure is braced on one side, leaving the other side open to form a larger room with another module. In general,
each module provides a single bedroom, whereas two open-sided modules are joined to create a larger room. Three modules are used to create
the lounge space (the middle module is open on both sides). For each typical floor, 25 modules are used. The various types of steel modules used
in the building are shown in Figure 2.

F I G U R E 2 Demarcation plan
for cast-in-situ structure and
modules
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
4 of 20 SHAN AND PAN

FIGURE 3 Conceptual design for steel module

2.3 | Preliminary design of structural components

The thicknesses of the RC core walls are assumed to be 450 and 250 mm for the interior and exterior core walls, respectively (Figure 2). The RC
coupling beams are assumed to have a section size of 400 × 800 mm (width × depth). The S355 steel with a yield strength of 355 MPa is adopted
for the structural steel of the modules. The combined dead and live loads distributed over a 7.1-m2 tributary are calculated as 10 kPa. The maxi-
mum axial load of the columns at the bottom story is 2,524 kN; accordingly, RHS 200 × 200 × 12.5 mm is adopted for the steel columns in the
modules, providing an axial load capacity of 2,567 kN. For the steel beam in each module, RHS 200 × 150 × 6 mm is used. A steel bracing with a
section size of 100 × 100 × 6.0 mm is added to the steel frame to improve module stability. The slenderness ratios of the steel columns and brac-
ings are 28.7 and 167.2, respectively; these do not exceed the allowable limit prescribed by the Hong Kong steel structure design code[24]
(i.e., 200 for structural compression members). The building density of the modular building is approximately 4.8 kN/m3, which is slightly less than
that of the reference building (5.0 kN/m3).

2.4 | Connections

Bolted connections are adopted to link the modules, because employing them can reduce the site work in comparison with the typical practice of
using welded joints. The three typical connections (i.e., module-to-module, module-to-podium, and module-to-core wall connections) used in
modular buildings are illustrated in Figure 4. In each typical connection, a row of two bolts is adopted. One 150 × 12 mm2 (width × thickness) tie
plate is set horizontally to connect two adjacent modules. Two high-strength 39-mm diameter steel anchor rods are employed to connect the
steel module and RC members (core wall and podium); the details of these connections can be found in a previous report.[25]

3 | N U M E R I C A L M O D E L I N G M E T HO D

To analyze the global behaviors of buildings under various loading scenarios, a three-dimensional (3D) numerical model is established using ETABS
software.[26] The local behaviors of modular connections are investigated via a detailed FE model using ABAQUS software.[27] The material prop-
erties of the concrete, reinforcing bars, and structural steel are summarized in Table 1.

3.1 | Reference building model

In the ETABS model, the frame members and shear walls are modeled using frame and shell elements, respectively. The floor slabs are modeled
using membrane elements that do not consider the out-of-plane stiffness; thus, no bending moment is transferred between the vertical structural
members and floor slabs. Moreover, the contribution of floor slabs to the stiffness of beams is neglected. A rigid diaphragm behavior is assumed
for all floors; accordingly, all nodal points are assumed to have no relative in-plane deformation. The gravity loads on floor slabs are assumed to
be distributed to vertical structural members according to the tributary area concept.[26] The connections between columns and foundations are
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SHAN AND PAN 5 of 20

F I G U R E 4 Schematic view of modular connections. (a) Module-to-module connection; (b) module-to-podium connection; (c) module-to-core
wall connection

modeled as pins. The modulus of elasticity of concrete is estimated according to the Hong Kong design code for concrete structures.[28] The
numerical model of the building is shown in Figure 5, where the red and blue contours indicate the shear wall and beam regions, respectively. The
figure also shows the dimensions and layout of structural components in the reference building.

3.2 | Modular building model

It is widely known that the implementation of a detailed FE analysis of modular buildings in the design process is extremely time-consum-
ing; accordingly, suitable simplifications are adopted. The horizontal connections between two adjacent modules are modeled by assigning a
link element (Figure 6a). The connection between the upper and lower modules is assumed to be pin-ended (Figure 6a) because the bolts
in the module-to-module connection may loosen as a result of vibrations during their service life. Moreover, because only one row of bolts
is adopted for this connection, an independent rotation would be generated between the upper and lower modules. Similarly, the pin-end

TABLE 1 Material properties of the concrete, reinforcing bars, and structural steel

Concrete Compressive strength 45 Mpa


Modulus of elasticity 26,400 Mpa
Density 2,450 kg/m3
Reinforcing bars Yield strength 500 MPa
Structural steel Yield strength 355 MPa
Tensile strength 510 MPa
Modulus of elasticity 205,000 MPa
Density 7,850 kg/m3
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
6 of 20 SHAN AND PAN

FIGURE 5 Numerical model of the RC building (units: m)

boundary condition is also adopted in the module-to-podium connections (Figure 6b). The connections between the modules and core walls
are simulated using two parts of steel rods. The first part is rigid to simulate the cast-in bolts (rigid part) installed in the cast in situ con-
crete. The second part starts from the center of the steel column and progresses to the surface of the RC beam/wall (connection part);
these parts are connected using the pin-end boundary condition (Figure 6c). It should be mentioned that in reality, a small bending moment
may be induced in each typical connection, and the simplification method adopted in this study may yield conservative results. However,
because this study primarily focuses on the structural design of modular high-rise buildings, conservative predictions are acceptable in the
design. The bracings in the steel modules are simulated using frame elements, and the connections between the bracing and modular frame
are also modeled by the pin-end boundary.
To avoid a convergence problem during the calculation process, the bottom layer of each module is not simulated in the FE model. Thus, no
double beam action is considered in each story, and conservative results may have been obtained. The entire modular building is illustrated in
Figure 7, and the typical floor system with modules in the ETABS model is illustrated in Figure 8. The partly connected diaphragms are adopted to
simulate the behavior of the floor system in the modular building. The initial imperfections of the steel members are not considered in the ana-
lyses; accordingly, it is recommended that more detailed and advanced analyses (e.g., considering initial imperfections) of modular buildings be
performed in future studies.

FIGURE 6 Schematic view of modeling method of modular connections in ETABS program


15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SHAN AND PAN 7 of 20

F I G U R E 7 Numerical model
layout of 31-story modular
building (units: m)

3.3 | Detailed FE model connections

To investigate the local behaviors of modular connections and ensure that all connections behave elastically under various critical loading scenar-
ios, detailed FE models are developed using ABAQUS software.[27] The 3D-stress solid element is used in the detailed FE model, and the mesh
sizes for different components are discreetly selected such that the model is convergent, and accurate numerical results are obtained. The FE
model with the module-to-module, module-to-core wall, and module-to-podium connection mesh sizes is depicted in Figures 9–11, respectively.
Elastoplastic material properties are assigned to the different steel components of modular connections: The modulus of elasticity and Poisson's
ratio are considered as 210 GPa and 0.3, respectively. The yield strengths of bolts and anchor rods are 1,100 and 690 MPa, respectively. The yield

FIGURE 8 3-D view of typical floor layout and modules of modular structure in ETABS model (units: m)
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
8 of 20 SHAN AND PAN

FIGURE 9 Finite element model of module-to-module connection (units: mm)

strength of steel frame members is 355 MPa. The tangential interaction behavior between two surfaces in contact is considered by the penalty
friction formulation with a friction coefficient of 0.2; the normal interaction behavior is considered using “hard” contact. The resultant structural
loads, including the column end moment, shear force, and vertical load, obtained from the global analysis using the ETABS model are applied to
the detailed FE model.

3.4 | Validation of numerical model

Detailed FE models are developed using ABAQUS software,[27] and 3D-stress solid elements[27] have been widely employed to simulate the local
behaviors of modular connections. Lacey et al.[29–31] simulated a series of modular connections using a modeling method similar to that employed
in the present study, and the numerical results are found to be in good agreement with test data.[29–31]
The ETABS software[26] has been successfully employed for the analysis and design of high-rise buildings.[13,14,16] In these previous studies,
similar modeling techniques, such as the use of frames for beams and columns and shell elements for shear walls, were adopted.[13,14,16] More-
over, the pin-end boundary condition has been widely used to simulate the connection between the upper and lower modules[8,9]; it has been
proven capable of predicting the experimental results of modular steel-braced frames under earthquake loading scenarios with satisfactory accu-
racy.[6] Moreover, the link element was successfully used by Fathieh and Mercan[9] to mimic the horizontal connection between two adjacent
modules.
The results of a modular steel-braced frame test performed by Annan et al.[6] were adopted to verify the modeling techniques of modular
connections in the present study. The details of the tested steel-braced modular frame specimen are presented in Figure 12. The yield strength of
the steel material was 480 MPa, which was determined via the 0.2% strain offset method.[6] The tested beams were W100 × 19 sections bending
about their strong axis with a width and depth of 103 and 106 mm, respectively; the flange and web thicknesses were 8.8 and 7.1 mm, respec-
tively. The columns were made of HSS51 × 51 × 5 square hollow structural sections (nominal dimension, 51 mm; wall thickness, 4.78 mm). The
braces were fabricated using HSS32 × 32 × 3 tube sections (nominal dimension, 32 mm; wall thickness, 3.18 mm). The tested modular steel-
braced frame was restrained to prevent out-of-plane movement. In the test, the upper module floor beam was loaded laterally until a 45-mm dis-
placement was reached. In Figure 13, the lateral load versus displacement obtained by the numerical model is compared with the experimental
and theoretical results of Annan et al.[6] It is observed that the numerical response is considerably consistent with the experimental results and
theoretical predictions in both elastic and inelastic stages.
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
9 of 20

Finite element model of module-to-podium connection (units: mm)


F I G U R E 1 0 Finite element
model of module-to-core wall
connection (units: mm)
SHAN AND PAN

FIGURE 11
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
10 of 20 SHAN AND PAN

F I G U R E 1 2 Tested modular steel-braced


frame (units: mm)[6]

4 | L O A D I N G SC E N A R I O S A N D SE I S M I C W A V E S E L E C T I O N

Static analysis is adopted to calculate the global responses of the reference and modular buildings under wind load according to the Hong Kong
design codes.[32] The elastic time history analysis and response spectrum method are adopted to compute the global responses of these buildings
under frequent earthquakes. Hong Kong has no seismic code; hence, the Chinese seismic design code[22] is utilized in this study. The anti-seismic
grade is Level 2, the seismic precautionary intensity is 7, and the design seismic group is 2.[22] The building is assumed to be located in a Class I1
site whose characteristic period is 0.3 s.[22] The seismic analysis is performed according to the following steps.
First, the seismic excitation waves and response spectrum (earthquake load) are inputted in the Y direction (short direction, Figure 1b).
The maximum responses of the two buildings obtained from the seismic excitation waves and response spectra are compared with those
obtained from the wind load. In the second step, two-dimensional earthquake loads are inputted in the Y (primary direction, Figure 1b) and
X (secondary direction, Figure 1b) directions to consider the simultaneous torsional effects on the buildings and obtain the internal forces
in the structural members. The ratio of the peak accelerations in the Y and X directions is assumed to be 1:0.85. The load combination 1.2
(DL + 0.5 × LL) + 1.3SL + 0.2(1.4WL) is used[22] in the analysis, where DL, LL, and WL are the dead, live, and wind loads, respectively; SL
denotes the maximum envelope of internal forces obtained via the seismic excitation analysis and response spectrum method. The buildings
are strong in the X direction, where the overall responses are significantly smaller than those in the Y direction. Accordingly, the structural
responses in the X direction are not presented.
The Loma Prieta wave, the Duzce wave, and an artificial wave are selected for seismic time history analyses. To represent the cases of fre-
quent earthquake, the peak acceleration is 50 cm/s2. The Newmark direct integral method is adopted in the dynamic analysis. After the seismic
waves are transformed to the seismic response spectrum, the difference between the accelerations of the seismic response and design response
spectra is found to be <20% in the critical natural period, as shown in Figure 14. Thus, the seismic waves adopted in this study are consistent with
the seismic wave selection criteria in the Chinese design code.[22]

5 | STRUCTURAL RESPONSE

5.1 | Modal analysis

To obtain the mode shapes, natural periods, and mass participation coefficients of the two buildings, the Wilson–Ritz vector method is adopted
for the modal analysis. A mass participation ratio of 95% is achieved after 18 mode shapes. Mode shapes 1–4 of the reference and modular build-
ings are shown in Figures 15 and 16, respectively. The first to the sixth natural periods of the reference building are T1 = 3.17 s, T2 = 2.88 s,
T3 = 2.32 s, T4 = 0.79 s, T5 = 0.7771 s, and T6 = 0.5929 s, respectively; those of the modular building are T1 = 2.05 s, T2 = 1.12 s, T3 = 1.04 s,
T4 = 0.49 s, T5 = 0.33 s, and T6 = 0.31 s, respectively. The natural periods of the modular building are smaller than those of the reference building
because of the higher lateral stiffness of the former. The natural periods of the reference and modular buildings approximate those of the other
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SHAN AND PAN 11 of 20

F I G U R E 1 3 Comparison of numerical, experimental, and


theoretical results of modular steel-braced frame[6]

F I G U R E 1 4 Comparisons between design response spectrum and


seismic response spectrum

F I G U R E 1 5 The first- to the fourth-mode shapes for the reference RC building (scale factor 500). (a) First order; (b) second order; (c) third
order; (d) fourth order

high-rise buildings investigated in previous studies.[16,33] Li et al.[33] performed a modal analysis of several 20-story coupled shear wall buildings
(72 m high) and reported that the first to the third natural periods of these buildings were approximately 2.145, 2.016, and 1.656 s, respectively.
Liu et al.[16] also reported that the first to the sixth natural periods of a 30-story modular prefabricated steel-framed high-rise building (99 m high)
were 3.29, 3.22, 2.45, 1.12, 1.1, and 0.8 s, respectively.
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
12 of 20 SHAN AND PAN

F I G U R E 1 6 The first- to the fourth-mode shapes for the 31-story modular building (scale factor 1,000). (a) First order; (b) second order;
(c) third order; (d) fourth order

5.2 | Comparison of global behaviors between wind and earthquake loads

The overall behaviors of the reference and modular buildings are compared under wind and earthquake loads. The results indicate that the
lateral displacements of the two buildings are practically linearly proportional to the building height. The maximum roof lateral displace-
ments under the wind load are approximately 121.0 and 36.2 mm for the reference and modular buildings, respectively. These values are
considerably less than the allowable limit set by the Hong Kong design code for concrete and steel structures.[24,28] The allowable elastic
roof lateral displacement under the wind load is 164 mm (height/500 = 82 m/500). The allowable elastic roof lateral displacements under
earthquake load are not prescribed in the Hong Kong or Chinese codes. It should be noted that the compound lateral deflection caused by
the flexibility of the transfer structure, elastic shortening of the columns, and lateral drift of the moment frame is not considered, resulting
in a smaller lateral drift.
The profiles of the interstory drift ratio, which is defined as the ratio of the interstory drift to the story height, are shown in
Figure 17. The maximum interstory drift ratios under the wind load were 0.056% (1.75 mm) and 0.016% (0.5 mm) for the reference RC
building and steel-framed modular building, respectively. These results are within the limits set by the Hong Kong code for wind load sce-
narios, that is, story height/400 (8.3 mm).[24] The maximum interstory drift ratios under the earthquake load are found to be 0.130%
(4.091 mm) and 0.047% (1.493 mm) for the reference and modular buildings, respectively. As shown in Figure 17, the interstory drift ratio
profiles of the two buildings under the earthquake load are significantly larger than those under the wind load. Therefore, conservative
results are obtained if the earthquake load is considered in the design of high-rise buildings in Hong Kong. The interstory drift ratio shapes
of the two buildings under earthquake loads significantly differ from those under the wind load. Under the earthquake load, these shapes
significantly change at the mid-height of the two buildings. This is because the global behaviors of high-rise buildings are considerably
affected by high-order vibration modes under earthquake excitations; similar observations were reported by Li et al.[33] and Liu et al.[16] In
their studies, the interstory drift ratio shapes are significantly altered at the mid-height of high-rise buildings under earthquake excitations.

F I G U R E 1 7 Comparison of
interstory drift ratio for reference
RC building and 31-story modular
building
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SHAN AND PAN 13 of 20

F I G U R E 1 8 Comparison of
story shear for reference RC
building and 31-story modular
building

F I G U R E 1 9 Influence of
interior core wall thickness on the
global behaviors of the 31-story
modular building under
combined load

F I G U R E 2 0 Influence of
exterior core wall thickness on
the global behaviors of the
31-story modular building under
combined load

The lateral displacements and interstory drift ratios of the modular building are significantly smaller than those of the reference building
because of the higher lateral stiffness of the former.
The maximum interstory drift ratio of the modular building under the combined load is 0.071% (2.238 mm). This is less than the allowable
elastic limit of the Chinese seismic code[22] in which the allowable elastic interstory drift ratio[22] is 0.125% (3.938 mm). The maximum interstory
drift ratio of the reference building is 0.251% (7.922 mm), which exceeds the allowable elastic limit of the Chinese code.[22] The reference RC
building is designed according to the Hong Kong wind code rather than the seismic code; thus, it poorly behaves during an earthquake. The lateral
displacement, interstory drift ratio, and story shear of the two buildings in the X direction are significantly smaller than those in the Y direction;
hence, they are not presented in detail in this study.
The story shear (Figure 18) is an important benchmark in building design and should be determined by evaluating the shear demand at
each story of a structure. The story shear demand of the modular building is approximately twice that of the reference building in the
earthquake loading scenario because of the higher lateral stiffness of the former. The story shears of the two buildings under the combined
loads are also shown in Figure 18. The internal forces in structural members (e.g., core walls, coupling beams, beams, and columns and
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
14 of 20 SHAN AND PAN

TABLE 2 Influence of interior wall thickness on the natural periods of modular building

Modal number

Wall thickness (mm) First Second Third Fourth Fifth Sixth


300–250 2.0616 1.161 1.0718 0.5154 0.3428 0.3158
450–250 2.0496 1.1200 1.0427 0.4903 0.3294 0.3064
600–250 2.0676 1.1067 1.0425 0.4811 0.3252 0.3033
750–250 2.0927 1.1073 1.0510 0.4779 0.3241 0.3032
900–250 2.1187 1.1150 1.0604 0.4771 0.3240 0.3044

Note: The first value in the wall thickness volume denotes the interior wall thickness, and the second value in the wall thickness volume denotes the exte-
rior wall thickness.

TABLE 3 Influence of exterior wall thickness on the natural periods of modular building

Modal number

Wall thickness (mm) First Second Third Fourth Fifth Sixth


450–100 2.1221 1.1355 1.0600 0.4980 0.3351 0.3078
450–250 2.0496 1.1200 1.0427 0.4903 0.3294 0.3064
450–400 1.9857 1.1171 1.0306 0.4812 0.3251 0.3075
450–550 1.9373 1.1193 1.0268 0.4723 0.3236 0.3121
450–700 1.9037 1.1252 1.0268 0.4723 0.3236 0.3121

Note: The first value in the wall thickness volume denotes the interior wall thickness, and the second value in the wall thickness volume denotes the exte-
rior wall thickness.

TABLE 4 A summary of critical connection forces in 31-story modular building

Shear force (kN)

Connection type Load type Axial force (kN) X direction Y direction


Module-to-module Max. shear −773.86 19.73 0
Max. tension 824.12 5.03 0
Module-to-core wall Max. shear 53.82 32.44 −221.40
Max. tension 99.92 6.00 −22.52
Module-to-podium Max. shear −678.66 0.10 1.55
Max. tension 848.25 0.11 0.23

Note: Negative value represents compression force, and positive value represents tension force in the column of axial force.

braces in steel modules) are within the allowable limits. Based on the shear distribution calculation, more than 90% of the story shear is
sustained by the RC core wall system, and less than 10% is borne by the steel modules. However, the results are not discussed in detail
herein because of the length limit of this paper.

5.3 | Influence of core wall thickness on global behaviors of modular buildings

The core wall thickness may vary depending on the design requirements. The effects of the interior and exterior core wall thicknesses on the roof
lateral displacement and interior drift ratio of the modular building under combined load scenarios with the other parameters kept constant are
investigated, respectively. The roof lateral displacements and interstory drift ratios gradually decrease and gradually increase with the increase in
interior (Figure 19) and exterior wall thicknesses (Figure 20), respectively. The fundamental period (first period) of the modular building gradually
increases with increasing interior wall thickness (Table 2). Therefore, a smaller earthquake load acts on a building with increasing interior wall
thickness. Furthermore, increasing the thickness of interior walls improves the robustness of the building and reduces the roof displacement and
interstory drift ratio. In contrast, the fundamental period of the modular building gradually decreases with increasing exterior wall thickness
(Table 3). Consequently, the earthquake load acting on the building increases, resulting in a larger roof displacement and interstory drift ratio. In
engineering practice, the trial-and-error method is recommended for determining the optimal core wall thickness.
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SHAN AND PAN 15 of 20

FIGURE 21 Stress contour of the critical connections for the 31-story modular building under the maximum shear scenario

5.4 | Connection behaviors

To determine the stress levels in the various components of connections under the worst loading conditions, the connections in the steel-framed
modules are simulated using the detailed FE modeling method. The internal end forces of steel-framed members surrounding the connections are
obtained from the ETABS model and then applied to the detailed FE models in ABAQUS software.[27] The performance of the modular connec-
tions in the ABAQUS model justifies the assumptions made in the ETABS model and provides valuable results for the local behaviors. The critical
connection forces obtained from the combined loads in the ETABS model are summarized in Table 4. The X and Y direction values listed in Table 4
are illustrated in Figures 9–11.
The stress contours of different connections in the modular building under critical connection forces are shown in Figures 21 and 22. It is evi-
dent that the maximum stresses in the connections are within the yield limits. Therefore, the design of the connections presented in this study is
satisfactory to resist the loads generated by frequent earthquakes.

5.5 | Load redistribution mechanisms

The redesigned modular building is a hybrid structure that adopts a steel-framed modular system connected to a core wall system. The system is
assembled using typical steel-framed modules with specially designed connections.
Vertical floor loads, including dead and live loads (DL and LL, respectively), should all be situated on the deck of each module and the RC slabs
in the core wall system. For the steel-framed modular system, the vertical floor loads are transferred from the slabs to their supporting secondary
steel beams and then to the primary steel beams. The loads on the primary beams are transmitted to the columns by the bending moment and
shear force; the columns eventually transmit the loads to the deep transfer slab. For the core wall system, the loads are transferred from the slabs
to the coupling RC beams and core walls. The loads from the coupling beams are transmitted to the walls by the bending moment and shear force.
The walls eventually transmit the loads to the deep transfer slab and podium.
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SHAN AND PAN

design solution for high-rise buildings using steel


F I G U R E 2 3 Workflow of the structural
Stress contour of the critical connections for the 31-story modular building under the maximum tension scenario

modules
FIGURE 22
16 of 20
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SHAN AND PAN 17 of 20

F I G U R E 2 4 Comparison of
global responses of 40-story
modular building under wind load,
response spectrum method,
seismic excitation, and
combined load

The lateral loads acting on the steel modular system, for example, wind and/or earthquake loads, are transferred through the portal frame to
the core wall system. The lateral loads acting on the core wall system are received by the coupling beams and core walls. Together with the RC
slab, the coupling beams provide a combined effect to achieve a rigid in-plane diaphragm for the core wall system, such that the lateral loads could
be effectively distributed and transferred to the podium by the core walls.

6 | H Y P O T H E S I Z ED E X A M P LE (4 0 - S T O R Y M O D U L A R B U I L D I N G )

A practical workflow for designing a high-rise building using steel-framed modules is presented in Figure 23. In the first step, the system scheme
for the modular building is proposed, followed by the preliminary design of structural members. Thereafter, the global performances of the build-
ing should be checked under various loading scenarios to ensure that the building behaves within the allowable limits of the design codes. Fur-
thermore, the detailed FE model of modular connections should be developed, and the critical reaction forces in these connections obtained from
the global analysis should be rigorously inputted into the FE models. The stresses in the modular connections should be carefully checked to
ensure that the connections behave within the yielding limit under adverse loading scenarios. The internal forces in structural components should
also be carefully checked to ensure that these components perform elastically under adverse loading scenarios. If the allowable limits are
exceeded, the structural designers are required to modify the modular system scheme or the dimensions of structural members and re-analyze
the building until the structure behaves within the allowable limits. This workflow allows the structural engineers to design a steel-framed modular
high-rise building based on the design codes without any of the structural members yielding.
The proposed workflow is implemented to a 40-story steel-framed modular building, which has the same floor plan and system scheme as
the 31-story modular building. Moreover, the thicknesses of the interior and exterior core walls are increased to 600 mm. The ETABS and detailed
FE models of the 40-story modular building are developed to examine the workflow proposed in this study. The global responses, such as the lat-
eral displacement, interstory drift ratio, and story shear of the 40-story modular building, are calculated under wind, earthquake, and combined
loads. It is found that the maximum roof lateral displacement under the wind load is 83.5 mm, which is within the allowable limit (220.5 mm) of
the Hong Kong wind code.[32] As indicated in Figure 24, the maximum envelopes of the interstory drift ratio and story shear of the 40-story mod-
ular building obtained under the earthquake load are significantly higher than those under the wind load. Under the combined load, the maximum
interstory drift ratio of this modular building is 0.11%, which is within the allowable limit of the Chinese code.[22] The critical reaction forces in the
connections obtained under the worst loading scenarios in the ETABS model are summarized in Table 5. The stresses in the modular connections

TABLE 5 A summary of critical connection forces in 40-story modular building

Shear force (kN)

Connection type Load type Axial force (kN) X direction Y direction


Module-to-module Max. shear −1065.06 20.57 0
Max. tension 1116.85 4.21 0
Module-to-core wall Max. shear 72.63 68.46 −251.39
Max. tension 79.74 25.32 −120.88
Module-to-podium Max. shear −833.26 0.24 6.05
Max. tension 1062.07 0.24 0.64
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SHAN AND PAN

Stress contour of the critical connections for the 40-story modular building under the maximum tension scenario
Stress contour of the critical connections for the 40-story modular building under the maximum shear scenario

FIGURE 26
FIGURE 25
18 of 20
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
SHAN AND PAN 19 of 20

in the 40-story modular building under critical reaction forces are presented in Figures 25 and 26. It is observed that the stresses in the modular
connections are within their yielding limits. In the lateral load versus story drift curves in Figure 13, it is observed that the 40-story modular build-
ing behaves elastically under wind and earthquake loads.

7 | CO NC LUSIO NS

A structural design solution is developed for high-rise buildings using steel-framed modules. This solution is demonstrated using a real 31-story
student hostel building and is further examined in a case where the proposed solution is applied to a hypothetical 40-story building. The global
behaviors of the buildings under wind and earthquake load scenarios are studied. The lateral displacement, interstory drift ratio, and story shear
of the reference and redesigned modular buildings are determined and compared. The stresses in the modular connections are checked to ensure
that they are within the yield limit under the worst loading scenarios. The key findings and conclusions of this study are as follows.

1. The study demonstrates the suitability of the proposed structural design solution for designing high-rise buildings using steel-framed modules.
The maximum roof lateral displacements of the reference and 31-story steel-framed modular buildings under wind load are 105 and 35.6 mm,
respectively, which are within the allowable limit (164 mm) of the Hong Kong wind code. Under combined loads, the maximum interstory drift
ratio of the modular building is 0.072%, which is lesser than the allowable limit prescribed by the Chinese seismic code (0.125%). However,
the reference building exhibits poor performance under the combined loads because it is only designed according to the wind code require-
ments; the seismic requirements are not considered. Additionally, the stresses in the modular connections are within the yielding limits of steel
materials.
2. The reference and redesigned modular buildings have larger lateral displacements, interstory drift ratios, and story shear under the earthquake
load in comparison with those under the wind load. Thus, it is recommended that the global performance of modular high-rise buildings against
earthquake load be verified in the structural design. Furthermore, the seismic behaviors of the reference and modular buildings are significantly
influenced by the high-order vibration modes.
3. The lateral stiffness of the 31-story modular building is higher than that of the reference building; therefore, the natural periods, lateral dis-
placement, and interstory drift ratio of the former are significantly smaller than those of the latter. Moreover, the story shear in the 31-story
modular building is significantly larger than that in the reference building.
4. The core wall thickness can have both positive and negative influences on the global behavior; thus, the trial-and-error method is rec-
ommended for determining the optimal core-wall thickness in engineering design.

Further research is necessary to improve the lateral stiffness of modular buildings, for example, by optimizing the structural layout and using
high-performance steel and high-strength concrete. Moreover, experiments should be conducted to verify the performance of critical structural
components and connections considered in this study.

ACKNOWLEDGEMEN TS
The authors acknowledge support from the Development Bureau of the Government of the Hong Kong Special Administrative Region (Project
No.: 200008191) and the Research Impact Fund of the Hong Kong Research Grants Council (Project No.: R7027-18). Also acknowledged are
access to the case building provided by the Estates Office of The University of Hong Kong and contribution of Ben Young, Yancheng Cai, and
Man-tai Chen to the project reported.

ORCID
Sidi Shan https://orcid.org/0000-0002-2593-3670
Wei Pan https://orcid.org/0000-0002-2720-3073

RE FE R ENC E S
[1] W. Pan, L. Chen, W. T. Zhan, ASCE J. Manag. Eng. 2019, 35(1), 05018013.
[2] C. I. Goodier, A. G. F. Gibb, Constr. Manag. Econ. 2007, 25(6), 585.
[3] W. Pan, A. G. F. Gibb, A. R. J. Dainty, Build. Res. Inf. 2008, 36(1), 56.
[4] W. Pan, C. K. Hon, Proc. Institut. Civil Eng. Mun.l Eng. 2018, 173, 64. https://doi.org/10.1680/jmuen.18.00028
[5] A. W. Lacey, W. S. Chen, H. Hao, K. M. Bi, J. Build. Eng. 2018, 16, 45.
[6] C. D. Annan, M. A. Youssef, EI Naggar M.H., Eng. Struct. 2009, 31, 1435.
[7] S.-G. Hong, B.-H. Cho, K.-S. Chung, J.-H. Moon, Eng. Struct. 2011, 67, 936.
[8] C. D. Annan, M. A. Youssef, EI Naggar M.H., J. Earthq. Eng. 2009, 13, 1065.
[9] A. Fathieh, O. Mercan, Eng. Struct. 2016, 122, 83.
15417808, 2020, 15, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/tal.1788 by chen haohou - Capital Normal University , Wiley Online Library on [21/08/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
20 of 20 SHAN AND PAN

[10] T. Gunawardena, T. D. Ngo, P. Mendis, L. Aye, J. Alfano, Structural performance under lateral loads of innovative prefabricated modular structures.
Materials to Structures: Advancement Through Innovation, Taylor & Francis, London 2013.
[11] S. Srisangeerthanan, M. Hashemi, P. Rajeev, E. Gad, Eng. Struct. 2018, 163, 25.
[12] X. Lu, B. Zhou, B. Zhao, W. Lu, Struct. Design Tall Spec. Build. 2015, 24, 1019.
[13] M. Shin, T. H.-K. Kang, J. M. LaFave, J. S. Grossman, Struct. Design Tall Spec. Build. 2012, 21, 918.
[14] A. Mwafy, S. Khalifa, Struct. Design Tall Spec. Build. 2017, 26, e1399.
[15] R.-Q. Feng, J. Ye, G. Yan, Q.-X. Li, B. Yao, Struct. Design Tall Spec. Build. 2013, 22, 802.
[16] X. Liu, X. Zhou, A. Zhang, C. Tian, X. Zhang, Y. Tan, Struct. Design Tall Spec. Build. 2018, 27, e1415.
[17] M. Lawson, R. Ogden, C. Goodier, Design in modular construction, CRC Press 2014.
[18] K. S. Choi, H. C. Lee, H. J. Kim, J. Korea Institut. Struct. Maint. Inspect. 2016, 20, 74.
[19] J. H. Doh, N. M. Ho, D. Miller, T. Peters, D. Carlson, J. Steel Struct. Construct. 2016, 2, 2472.
[20] T. Gunawardena, Behavior of prefabricated modular buildings subjected to lateral loads, Doctor of Philosophy, The University of Melbourne, Melbourne,
Australia 2016.
[21] Z. Chen, J. Liu, Y. Yu, Eng. Struct. 2017, 147, 625.
[22] GB50011–2010, Code for seismic design of building, Ministry of Housing and Urban Rural Development, Beijing, China 2010.
[23] BD, Code of Practice for Dead and Imposed Loads, Building Department, Hong Kong 2011.
[24] BD, Code of Practice for the Structural Use of Steel, Building Department, Hong Kong 2011.
[25] S. Shan, D. Looi, Y. Cai, P. Ma, M.-T. Chen, R. Su, B. Young, W. Pan, Proc. Institut. Civil Eng. Civil Eng. 172, 51. https://doi.org/10.1680/jcien.18.00052
[26] ETABS. [00Computer software]. Computers and Structures, Inc.,Berkeley, CA. 2009.
[27] ABAQUS. [00Computer software]. Hibbitt, Karlsson, and Sorenson, Inc., Pawtucket, RI, USA. 2010
[28] BD, Code of Practice for Structural Use of Concrete, Building Department, Hong Kong 2013.
[29] A. Lacey, W. Chen, H. Hao, K. Bi, Eng. Struct. 2019, 198, 109465. https://doi.org/10.1016/j.engstruct.2019.109465
[30] A. Lacey, W. Chen, H. Hao, K. Bi, J. Constr. Steel Res. 2019, 158, 576. https://doi.org/10.1016/j.jcsr.2019.04.012
[31] A. Lacey, W. Chen, H. Hao, K. Bi, J. Constr. Steel Res. 2019, 162, 105707. https://doi.org/10.1016/j.jcsr.2019.105707
[32] BD, Code of Practice on Wind Effects in Hong Kong, Buildings Department, Hong Kong 2004.
[33] G.-Q. Li, M. Pang, F. Sun, J. Jiang, D. Hu, Struct. Design Tall Spec. Build. 2018, 27, e1405.

How to cite this article: Shan S, Pan W. Structural design of high-rise buildings using steel-framed modules: A case study in Hong Kong.
Struct Design Tall Spec Build. 2020;29:e1788. https://doi.org/10.1002/tal.1788

You might also like