You are on page 1of 24

Journal of Building Engineering 74 (2023) 106828

Contents lists available at ScienceDirect

Journal of Building Engineering


journal homepage: www.elsevier.com/locate/jobe

A technical review of computational fluid dynamics (CFD)


applications on wind design of tall buildings and structures: Past,
present and future
Kasun Wijesooriya a, Damith Mohotti a, *, Chi-King Lee a, Priyan Mendis b
a
School of Engineering and Information Technology, The University of New South Wales, Canberra, ACT, 2600, Australia
b
Department of Infrastructure Engineering, The University of Melbourne, Melbourne, 3010, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: Over the past two decades, an upsurge in using Computational Fluid Dynamics (CFD) for wind
Super-tall buildings design on tall buildings has been observed. An extensive amount of work has been performed,
Tall building where validation has been at the forefront of most of these studies. Challenges associated with
Fluid-structure interaction CFD and different methodologies used in the analysis haven’t been well articulated within the
Wind loading scope of tall building design. Also, there is a lack of critical best practice guidelines in using CFD
Computational fluid dynamics (CFD)
for wind analysis of structures which can be readily adopted by practitioners and researchers
Building design
alike. To this end, this paper presents a comprehensive technical review of the application of
Computational Fluid Dynamics (CFD) on tall buildings and structures. A thorough discussion of
CFD and its concerning design challenges specific to tall building design, such as turbulence
modelling, inflow turbulence and domain & mesh configurations, are discussed in detail.
Furthermore, a comprehensive literature of CFD studies on tall buildings is presented, covering all
important topics from basic bluff body aerodynamics to complex Fluid-Structure Interaction (FSI)
studies. Where applicable, the literature is critically evaluated (impartially) in this paper and
compared with supportive arguments from the author’s extensive experience. Finally, the
manuscript concludes with potential upcoming numerical methods such as the Lattice-Boltzmann
Method (LBM) and the use of Artificial Intelligence (AI) to design tall buildings.

1. Introduction
Modern super-tall buildings are susceptible to wind-induced motions, usually due to their slenderness and height. With recent
advancements in construction and material technology, super-tall structures are designed with strong, lightweight materials,
enhancing the implications of wind-induced motions. Structural serviceability is one of the main concerns of slender super-tall
structures as they experience low frequency and low acceleration building vibrations, affecting occupant comfort [1,2]. Further­
more, wind is the governing lateral load for which the structural integrity needs to be designed [2].

1.1. Experimental design techniques – a brief background


Wind tunnel testing is the industrial standard for evaluating wind-sensitive structures such as super-tall buildings. Wind tunnel
testing has evolved in the last century, where procedures for conducting tests in a scaled environment were presented and constantly

* Corresponding author. School of Engineering and Information Technology, The University of New South Wales, Canberra, 2600, ACT, Australia.
E-mail address: d.mohotti@unsw.edu.au (D. Mohotti).

https://doi.org/10.1016/j.jobe.2023.106828
Received 12 January 2023; Received in revised form 3 May 2023; Accepted 10 May 2023
Available online 19 May 2023
2352-7102/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

evaluated to improve the technique further. Many researchers have made significant contributions over the years to evolve wind
tunnel techniques. However, Jensen’s study [3] is perhaps the most pivotal. Jensen [3] introduced the importance of scaling laws and
the need to conduct experiments under boundary layer conditions. Since then, wind tunnel testing in terms of testing techniques,
instrumentation and acquisition methods has evolved, all of which have been comprehensively detailed and documented by Leighton
& Cochran [4].
In modern wind tunnel practice, three prominent techniques are employed in the assessment of wind-sensitive buildings: (1) High-
Frequency Base Balance (HFBB) method [5], (2) High-Frequency Pressure Integration (HFPI) method [6] and the more intricate and
traditional (3) aeroelastic modelling method (which in some cases can also measure surface pressures as presented by Pozzuoli et al.
[7]). The HFBB and HFPI techniques are the most utilised wind tunnel techniques, especially to measure aerodynamic forces and
wind-induced pressure on the surface, respectively. The HFBB technique uses a spectral approach where the generalised forces on the
building are determined as the building mode shape is assumed to be linear [5]. The main advantage of the HFBB technique is its
capability to record exact forces and moments at the base. However it cannot determine cladding pressures. Since its first imple­
mentation, the HFBB technique has undergone a series of developments where mode shape corrections [8,9] and the influence of
higher-order modes [10–12] can be included. Even with these advancements, the HFBB technique poses difficulty at the model
fabrication stage. HFBB models need to be light and stiff, which is difficult to replicate for slender super-tall models. The inability to
produce such a model leads to resonant effects of the HFBB setup, which can corrupt the data of the acquired forces at the base [13].
High wind speeds can further introduce resonance effects, limiting the test setup to lower wind speeds where the Reynolds number
difference can affect the overall results [2]. Furthermore, for super-tall structures, the influence of higher modes is often found to be
critical and including coupled modes of vibration, which cannot be neglected. Thus, with the limitations of the HFBB technique, it is
not suitable to be used in the design of slender super-tall structures [2,6,14]. The HFPI method on the other-hand can measure cladding
pressures and vertical profile but are unable to measure base moments and forces to the same accuracy of HFBB (accuracy increases
with resolution of pressure taps). If sufficient pressure tap distribution is provided, the HFPI technique can predict responses for all
modes of vibration, leading to accurate predictions of structural response [15,16]. The HFPI model must be rigid and there are no
constraints to its mass (or model flexibility), making it an easier model to fabricate in terms of requirements. The accuracy of the HFPI
technique directly relates to capability of measuring higher resolution surface pressure maps and often this is limited by instrumen­
tation and surface intricacies especially of tall slender structures. Successful applications of the HFPI technique can be found in recent
literature where the structure’s response is determined via a modal analysis [15,16]. In recent literature there has been a shift towards
the use of multi-degree of freedom (MDOF) aeroelastic models for the design of slender super-tall structures [17–20]. The aeroelastic
method is the most comprehensive experimental procedure adopted in current wind tunnel design as it is capable of measuring all
defined structural responses [21]. A multi-degree of freedom (MDOF) aeroelastic model should be able to replicate all modes of vi­
bration with significant modal participation factors critical for the structure’s response. Considering the requirements and accurate
representation of the structure, the aeroelastic model is both challenging and costly to fabricate, especially in scenarios where
complicated vibration modes must be considered. For this reason, the aeroelastic model is often reserved for complex structures and is
usually in the final stage of the design [2]. The stick aeroelastic model is an alternative to an MDOF aeroelastic model. It is a single
degree of freedom model (SDOF) and is designed such that it only recreates the two fundamental sway modes of vibration, leading to
savings in time and costs [22]. The stick aeroelastic model is only suitable for structures with dominant fundamental sway modes
(regular geometry buildings of moderate height). Even then, it has been found to generate unrealistically larger building responses
compared to MDOF models [23].

1.2. The potential of numerical modelling using computational fluid dynamics (CFD)
With an increasing research interest in slender super-tall structures, the demand for MDOF aeroelastic models has been on an
upsurge. The creation of aeroelastic models is intricate and often leads to fabrication difficulties at the model scale, especially when
higher-order modes become significant and needs to be included. This can be a limiting factor when considering experimental models,
especially as buildings tend to be more complex. With recent advancements in computer hardware and technology, researchers and
designers have explored the possibility of using numerical modelling as an alternative to make predictions of wind-induced loads on
super-tall structures. Computational Fluid Dynamics (CFD) has been at the forefront of numerical modelling and commonly adopts the
Finite Volume Method (FVM) for discretisation in most commercial codes such as FLUENT [24]. The use of computational and nu­
merical methods in wind engineering problems is known as Computational Wind Engineering (CWE) [25], and CFD has been the major
contributor in this aspect. The study of CFD on tall and super-tall structures is only a sub-portion of a broader effort in CWE and covers
a wide scope of problems. The implementation of CFD within the field of CWE is increasingly gaining popularity mainly due to its
versatility and ability to predict flow phenomena computationally. As mentioned by Blocken [26] in 2014, CWE has successfully
transitioned from an emerging to an established field in wind engineering due to recent progress made in the research community.
Topics such as pedestrian level wind (PLW) [27–29], building comfort analysis [30,31] and pollutant dispersion [32–34] studies in
urban environments are already well established and shown to provide good numerical accuracy with the use of CFD. With the
improvement of computational hardware and turbulence modelling, the computational effort required has drastically reduced over the
recent past. Perhaps, the biggest advantage of such an approach is the ability to perform multiple parametric studies and the ability to
measure data within a spatial volume without being limited to instrumentation such as that encountered in an experimental setting.
Some of the earliest implementations of CFD on structures were studied on regular shapes, where external aerodynamics and flow
fields around the bluff bodies were studied [35]. These early studies were performed for 2D cross-sections where validations in terms of
stagnation points, pressure quantities and velocity fields (or vectors) and their distribution at its near vicinity and far regions were
measured. Validation studies were at the forefront of these initial implementations of CFD, where comparisons with experimental

2
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

outcomes were drawn with a focus on the applicability of different turbulence models for reproducing flow phenomena.
Scale-resolving turbulence models such as Large Eddy Simulations (LES) were deemed too expensive, and the use of Reynolds
Averaging Navier Stokes (RANS) models was pursued. Limitations on computer hardware ensured the development of various RANS
turbulence models that could only provide statistical quantities and did not resolve stochastic turbulent motions of flow. A good
account of these early studies and their outcomes are articulated in the early works of Rodi [36] and Murakami [35], where open flows
for bluff bodies are concerned. One of the important conclusions drawn by Rodi [36] at the time was that RANS turbulence models,
regardless of accuracy, would be critical in engineering applications due to their practicality. Also, both Rodi [36] and Murakami [35]
discussed the importance of LES simulations and their capability to produce accurate simulations in terms of both statistical and
transient flow features. However, at the time, they concluded that LES simulations, albeit being accurate, are computationally
expensive for large flow problems where high Reynolds number flows are encountered. Rodi [36] mentions that LES would be widely
used as computational hardware improves in the “foreseeable but more distant future”. Over the last four decades, following on from
Rodi [36] and Murakami [35], a wide array of research has been performed within the scope of estimating wind loads and flow
phenomena for bluff bodies. Studies conducted by Lübcke et al. [37], Rodi [38], Kravchenko [39] and Catalano et al. [40] are some
examples where the development of CFD techniques have been studied on bluff body aerodynamics such as square and circular
cylinders. More importantly, the rapid growth of various turbulence modelling techniques was observed. The k − omega shear stress
transport (SSTkω) [41,42] RANS turbulence model is a notable example which was formulated to accurately predict wall bounded
flows, a deficiency observed in k − ε type models. Also, hybrid RANS-LES models such as the Detached Eddy Simulation (DES) [43–46]
and Delayed-DES (DDES) were implemented to model scale resolving properties of flow at a lower computation cost.
The CFD numerical studies performed on bluff body aerodynamics played a pivotal role where most of the findings were translated
for civil building structures. Some of the early applications of CFD on buildings were mainly focused on low-rise structures of simple
geometry, such as the Texas Tech University (TTU) Building [47–49]. These early adaptations of CFD on low-rise structures and their
findings naturally extended to modelling on tall structures. The first and perhaps the most influential study of CFD on tall buildings was
performed by Huang et al. [50] in 2007. The analysis was performed on the standard CAARC building (Wardlaw & Moss [51]), where
unsteady-RANS (URANS) models and an LES model with synthetic turbulence were studied. Issues pertaining to the deficiency of RANS
models, especially in its context to tall buildings through measuring surface pressure and flow streamlines, were identified through the
study. More importantly, the influence of velocity and turbulence intensity profiles on mean and standard deviation pressure was
noted. Also, they demonstrated that LES simulations could provide reasonable estimates of wind-induced quantities but concluded that
more research was necessary.
Huang et al.’s [50] study paved the way for almost two decades of CFD research on tall buildings. Since then, researchers have
studied various topics pertaining to tall building design such as; the effects of shielding, corner modifications, building shape modi­
fications, where most of these were performed as parametric validation studies. More importantly, investigations of the applicability of
various turbulence models, grid generation techniques and inflow turbulence methods have also been studied over the years with a
focus on tall building design. Recent literature reviews on tall buildings can be found by Hou & Jafari [52], Jafari & Alice [53],
Abu-zidan et al. [54] and Thordal et al. [55], where they cover important aspects of tall building design both experimental and nu­
merical approaches. However, none of these literatures fully covers the evolution of CFD or the modelling challenges faced by CFD
with respect to tall buildings. Thus, this manuscript aims to encapsulate the research efforts made in the past few decades to showcase
the progress and notable achievements made in applying CFD on tall buildings, including future prospects. Also, the challenges faced
by current modelling techniques will be discussed, such as Reynolds number effects, importance of turbulence modelling, generating
inflow profiles, sub-grid scale modelling, mesh generation techniques, etc. More importantly, the ability to provide direct structural
response measurements through current state of the art fluid-structure interaction (FSI) methods is also discussed. Where possible, the
outcomes from different researchers will be critically evaluated and presented with supportive arguments. Finally, every effort has
been taken to provide a comprehensive and fair outlook of all studies related to CFD applications on tall buildings based on the
extensive experience of the authors in using CFD on tall and super-tall building design.

2. numerical modelling and its challenges


This section presents the critical numerical modelling considerations required in the application of CFD on tall buildings. The
choice of turbulence models, inflow turbulence generation, domain and mesh creation are discussed. These four key parameters are
considered to be at the heart of CFD numerical modelling for tall buildings, and this section delves through the numerical modelling
considerations and challenges associated with them. Thus, this section serves as complete literature regarding the evolution of these
critical aspects with some recommendations. The final aim is to provide the reader with the underlying considerations that need to be
taken before choosing an appropriate model for the problem at hand.

2.1. Turbulence modelling


In Computational Fluid Dynamics (CFD) for a particular flow problem, the finite element (FE) code is solved via the Navier-Stokes
(N–S) equations within a spatially discretised volume. Turbulence Models are used to describe the flow of the fluid within an enclosed
domain, and the level of complexity of the modelled flow depends on how the turbulence model solves the N–S equations. For most
engineering applications where high Reynolds (Re) number flows are considered, directly solving the N–S equations is unfeasible, as
resolving the entire spectrum of turbulence length scales takes considerable computational effort. Such a simulation is often referred to
as Direct Numerical Simulation (DNS) and the number of grid points required usually scales as a function of Re9/4 and generally takes a
CPU time of Re3 [56]. For context, Huang et al.’s [50] study of the CAARC building (where Re > 105 ) would roughly require 180

3
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

billion grid cells to effectively perform a DNS simulation to capture all scales of turbulence. It is evident that DNS is unfeasible within
the scope of tall building design, even at the model scale. Hence in general, turbulence is often modelled, and the degree by which
turbulent scales are resolved depends on the complexity of the turbulence model employed. Thus a hierarchy of turbulence models can
be assorted where DNS followed by LES is at the very top in terms of complexity and 1st order RANS turbulence models (which model
all scales of flow) lie at the bottom [56]. Fig. 1 presents the complexities of each turbulence model to solve the N–S equations by
comparing the power spectral density (PSD) of a wavelength as illustrated in the work of Thordal et al. [55]. The distinction between
DNS, LES and RANS models can be clearly highlighted and understood in Fig. 1. It must be noted however that applications of DNS are
becoming increasingly prevalent in recent literature. For instance, Diaz-Daniel et al. [57] conducted 3D-DNS simulations for a cube
immersed in laminar and turbulent boundary layers with a maximum Reynolds number of 3000 (based on cube height and freestream
velocity). Their study focused on flow structures generated downstream and were compared with experimental outcomes. Chen et al.
[58] studied the influence a mounted circular cylinder near a stationary wall at subcritical Reynolds number of 500 using 3D-DNS. In
another study, Chen et al. [59] investigated the single degree of freedom FSI interference effects of two circular cylinders with various
gap ratios. They used 3D-DNS simulations to describe flow features on the observed cylinder oscillations. Finally, Zhao et al. [60]
studied interference effects between two rectangular cylinders at different width ratios using 3D-DNS simulations at a maximum
Reynolds number of 1000. All these studies were mostly focused on investigating fundamental flow physics, where 3D-DNS can reveal
better insight into important scale features. However, it must be noted that almost all these cited 3D-DNS simulations are investigated
at smaller Reynolds numbers, in smaller domains, compared to what is required for tall building design as previously discussed.
The choice of turbulence model for tall building applications needs to be selected appropriately depending on the required out­
comes. RANS turbulence models are generally used as a preliminary analysis to obtain time-averaged solutions, such as mean pressure
quantities, at a quick turnaround. Huang et al. [50] evaluated the performance of the CAARC building for three different variants of the
of k − ε (kε) Unsteady-RANS (URANS) models; (1) Standard-kε (Skε), (2) the Launder-Kato (LK kε) and (3) Murakami, Mochida and
Kondo MMK kε. It was shown that the MMK kε could provide the closest estimates of statistical quantities of mean pressure, drag & lift
coefficients and flow streamlines. A common problem the kε type models faces is the over-prediction of kinetic energy at impinging
regions. The MMK modification alleviates the production of turbulent kinetic energy (TKE) via modifications of the eddy viscosity
term in the original formulation of k − ε model [61]. Meng et al. [62] performed a sensitivity analysis of the CAARC building for
Steady-RANS models available in commercial CFD codes. They showcased that the SST kω and RNG kε models provided the closest
estimates in terms of mean surface pressures. Tominaga [63] performed URANS simulations on a tall building (rectangular cylinder)
where SSTkω, Skε, RNG kε, and Realisable kε (Rkε) turbulence models were implemented to study large-scale formations of eddies. It
was shown that the SSTkω was capable of reproducing the unsteady fluctuation behind the building. However, the SSTkω models
under-predicted turbulent kinetic energy (TKE), which led to the over-prediction of separation at the building corner. Tominaga [63]
proposed a modification to the Epsilon term in the RNG kε, where sufficient TKE is reproduced comparable to experimental outcomes.
Despite these changes, it was still found that the URANS simulation still overestimated the reattachment lengths on the roof and behind
the building. The authors (Mohotti et al.) [64] presented a study on a super-tall building of unorthodox geometry with circular
cross-sections where Steady-RANS (SRANS) turbulence models were used. The circular nature of the geometry meant that flow sep­
aration occurred due to an adverse pressure gradient as opposed to separation at leading edges as observed in square/rectangular
shaped buildings. To determine the correct point of flow separation, it was shown that the SSTkω RANS model was the most capable.
This was achieved by coupling the SSTkω with a stringent mesh consisting a first cell non-dimensional height (y+ ) that conforms to;
y+ < 1. Despite the ability of SSTkω, it was observed that leeward pressures were poorly predicted and required a scale resolving

Fig. 1. Correlation between PSD and turbulence modelling [55].

4
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

turbulence model [64]. The non-dimensionless wall distance y+ is commonly described as a function of wall friction (uτ ), kinematic
viscosity (υ) and distance from the wall (yp ) as given in Eq. (1).
uτ yp
y+ = Eq. 1
υ
In general the choice of y+ depends on the turbulence model and on the flow problem at hand. For instance, as mentioned earlier,
the SSTkω model is usually opted for curved surfaces where point of flow separation (transition point) predictions is required and
hence y+ < 1 needs to be employed. For highly turbulent external flows the choice for RANS modelling, the Standard-kε (Skε) can be
opted where y+ > 30 with scalable wall functions is opted. Thus, the choice of y+ and turbulence model can be said to be inter-related
and needs to be selected in tandem to one another.
Large Eddy Simulation (LES) turbulence model resolves all scales of flow up to the inertial sub-range and models the viscous eddies
through sub-grid scale (SGS) modelling. From a numerical modelling perspective, LES resolves all scales of turbulent flows to the size of
the mesh [65]. Thus, the mesh size, especially in regions around the building surface, needs to be carefully employed to provide
sufficient resolution. The general consensus for the application of LES for tall buildings is to provide a first cell height with y+ < 1 at the
boundary with sufficient inflation layers and an appropriate growth rate. Naturally, the computational cost directly correlates to the y+
employed at the wall boundary of the subject structure. For structures with sharp edges, y+ values can be more lenient as opposed to
those with curved surfaces. This requirement for y+ < 1 is vital in cases with curved surfaces as the point of flow separation due to
adverse pressure gradient needs to be accurately determined. Thus the stringent mesh requirement naturally leads to high compu­
tational costs and is one of the main challenges with regard to using LES turbulence models. The fundamental basis of LES modelling is
based on the filtering of the N–S equations where all scales smaller than the filter width are modelled (SGS). Thus, the SGS model plays
an important role, especially at the boundary where viscous eddies are concerned. One of the basic SGS models often available in
commercial codes is the Smagorinsky-Lilly SGS model [66] (S − model), where the eddy viscosity is modelled by assuming a singular
model constant (Cs ). An extension to this model was provided by Germano et al. [67] and Lilly [68], who noted that a single universal
constant could not be used for LES modelling and, instead, Cs is calculated dynamically (D − model). A further adaptation of the D−
model is the dynamic kinetic energy model (Dksgs − model) proposed by Kim and Menon [69], which further improved the dynamic
model by accounting for turbulent kinetic energy. Initial implementations of LES on bluff bodies by Rodi [36,38] employed the S −
model. However, the use of a singular constant to evaluate eddy viscosity stresses was proven to be ineffective by Murakami [70].
Differing Cs values were required to accurately replicate various flow features [70]. Every situation demanded a different model
constant such as: for the approaching flow, recirculation of flow, and downstream free shear flow [70]. Given that tall buildings are
subjected to complex flow phenomena, using a single constant for resolving SGS stresses was unsuitable. The application of dynamic
models for tall buildings can be found in recent literature where the CAARC building [50,71,72] or square/rectangular prisms are
considered [70,73]. However, in a comprehensive study done by the authors (Wijesooriya et al.) [74], where five SGS models are
compared, the Wall Adapting Local Eddy (WALE) SGS model was found to be the most suitable. For a super-tall structure with circular
cross-sections, it was shown that the WALE model was able to predict the point of flow separation to the highest degree of accuracy.
The stochastic nature of wind coupled with the dynamic sensitivity of tall structures calls for LES type modelling capable of
simulating transient wind behaviour. On the other hand, RANS turbulence models are incapable of producing the same. Huang et al.
[50] compared the power spectral density (PSD) curves of crosswind forces for LES and MMK kε against experimental outcomes
obtained from Obasaju [75], as displayed in Fig. 2. From the PSD curve, the RANS MMK kε can only estimate the peak frequency of the
crosswind force, where a similar observation was reported by Tominaga [63]. In comparison, the LES model can replicate the PSD
curve, to a higher degree of accuracy, over all frequencies. It is clear from observing Fig. 2 LES model is far capable of replicating the
fluctuating forces of the structure due to its ability to resolve the scales of flow for a wider energy spectrum compared to RANS models.
The discrepancies observed between the PSD curve for LES and experimental could be attributed to the inflow turbulence generator, an
important modelling requirement in CFD simulations for tall buildings.

2.2. Inflow turbulence generation


One of the critical challenges of CFD modelling is the simulation of atmospheric boundary layer conditions within the numerical
domain. Much like experimental wind tunnel testing, the numerical domain needs to be capable of simulating the wind profile, tur­
bulence intensity profile, power spectra, spatial coherence and spatial correlation statistics [76]. There are various techniques

Fig. 2. Crosswind PSD comparison of CAARC building between; LES, MMK kε and Experimental.

5
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

available for generating inflow turbulence as outlined by Thordal et al. [55]: (a) Replication of wind tunnel, (b) Domain recycling
method, (c) Precursor domain method and (d) Synthetic turbulence, which are depicted from Fig. 3(a)–(d) respectively. Of these four
methods, the latter synthetic turbulence inflow method is the most preferred for a few reasons. Methods (a), (b) and (c) require more
computational resources due to the size of the domain and complexity of the flow, which is generated from the obstacles placed
(usually in the form of blocks) to create turbulence. The generated turbulence is highly sensitive to the mesh density employed at the
obstacles. Given that the blocks are generally smaller than the target building, a stringent mesh density may need to be applied to
sufficiently resolve the scales down to the inertial subrange. Also, due to the nature of the flow and its complexity, the ability of the
simulation to achieve convergence within a time step may be difficult and may require a more robust pressure-velocity coupling al­
gorithm to be employed. An example of wind tunnel domain replication and precursor domain application can be found in Thordal
et al. [77], where domain mesh sizes consisting of more than 30 million elements are observed.
Synthetic turbulence methods generate inflow turbulence by defining an algorithm that can provide random fluctuations. The most
common form of implementing synthetic turbulence in wind engineering applications is through the Fourier Transform-based
approach, such as the random flow generation (RFG) by Smirnov et al. [78], which is an adaptation of Kraichnan [79]. The RFG
was one of the first synthetic turbulence methods employed on tall building applications [50]. Still, as Huang et al. [80] discussed, the
RFG cannot reproduce fluctuations that provide a similar PSD to that of Von-Karman, which is encountered in an atmospheric
boundary layer (ABL). The RFG is based on the Gaussian model, and understanding this deficiency, Huang et al. [80] proposed the
Discretising Synthesising Random Flow Generator (DSRFG), which is an adaptation of the works [78,79] but imposing the Von-Karman
spectra instead of the Gaussian model.
Various models have been presented in literature, such as those by Castro et al. [81] and Kim et al. [82], which were further
adaptations of the DSRFG. However, recent literature has widely used, Aboshosha et al.’s [83] Consistent Discrete-RFG (CDRFG)
method. The CDRFG method is a further adaptation of the DSRFG, where Aboshosha et al. [83] corrected deficiencies in the target
velocity spectrum and spatial coherency function of velocities. Comparisons of alongwind, crosswind and torsional moments for the
CAARC building using the DSRFG and CDRFG are compared with wind tunnel results in Fig. 4. It can be seen that the CDRFG, in
particular, can represent moment plots closest to the experimental results. The use of CDRFG has been well received in the wind
engineering community, where many examples of its applications on tall buildings can be found in literature [84,85]. The CDRFG
method has also found its uses in CWE applications such as pedestrian level wind [86], investigation of ABL developments [87],
topographical effects [88] etc. In recent literature, new synthetic turbulence models based on the CDRFG can be found, such as in
Patruno and Ricci’s [89] (Prescribed-wavevector RFG (PRFG)) and Yu et al.’s [90] Narrow Band Synthesis RFG (NSRFG). Both PRFG
and NSRFG synthetic turbulence models aimed to improve the deficiencies of flow anisotropy and spatial length scale homogeneity
encountered in the CDRFG method.
Most studies often measure profiles, velocity spectra and spatial correlations at the inlet of the domain. Much like in an experi­
mental wind tunnel study, the approach flow velocity statistics must be measured at the building location. In CFD simulations, it is
often found that maintaining inflow conditions to the target location is difficult due to dissipation which is often a result of the
capability of the synthetic model employed [91]. The implication of measuring these quantities at the inlet as opposed to in an empty
domain at the target building location is highlighted in Section 3.1. Inadequate mesh sizing can also significantly influence the inflow
turbulent perturbations as it approaches the target building location. This was highlighted in the author’s (Wijesooriya et al.’s) [74]
study, where the vortex method was used to synthesise turbulent fluctuations. The study showed that the inflow characteristics of flow
are highly sensitive to grid size and that multiple parametric studies within an empty domain are required to adjust the inflow
characteristics. This indicates that some form of mesh refinement is required in the region from the inlet to the building location to
maintain the flow’s inflow characteristics. Thus, mesh refinement and placement of refined regions must be strategically employed
where both inflow characteristics and building induced velocity fluctuations are considered. The challenges of mesh refinement,
domain sizing and its direct influence on the quality of the simulation are discussed in the next section.

Fig. 3. Inflow turbulence generation methods [55]: (a) wind tunnel replication model, (b) precursor database method, (c) recycling method and, (d) synthetic tur­
bulence generator.

6
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

Fig. 4. Base moments spectra comparison between CDRFG and DSRFG [83].

2.3. Domain and mesh configurations


The CFD simulations’ domain is essentially the representation of a “wind tunnel” but within the numerical setting. The most
common setup for tall building applications is to enclose the subject with a six-sided rectangular domain where definitions of inlet,
outflow, walls and symmetries are defined, as shown in Fig. 5. At the inlet, velocity and turbulence profiles are defined. The input of the
velocity profile is relatively straight forward where the user defines a profile based on the applicable national wind code. However,
turbulence is usually defined as profiles of turbulent kinetic energy (TKE), turbulent dissipation rate (TDR) or specific dissipation rate
(SDR) which are given as follows:
( )2
TKE = α Uavg Iu (z) Eq. 2

TKE3/2
TDR = Cμ3/4 Eq. 3
L
ε
SDR = Eq. 4
kCμ

where Uavg is the average velocity, Iu (z) is the turbulence intensity (z is the vertical height), Cμ is a model constant with value of 0.09,
and L is the integral length scale. The integral length scale is obtained from the wind tunnel experimentations. In numerical simulations
of CFD, much like in the wind tunnel, the velocity and turbulent intensity profiles at the building location are ideally measured within

Fig. 5. Typical domain configuration for an isolated building with boundary definitions of inlet, outflow, side & top symmetries, ground wall and, building wall.

7
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

an empty domain. Thus, fine adjustments to the inlet TKE and TDR (or SDR) can be made to achieve the target profiles. Usually, the
quantity α is a choice between 0.5, 1.0 and 1.5 where in some literatures such as Tominaga et al. [92] and Blocken [93] a value of 1 is
recommended and Yan & Li [72] uses 1.5. However, as mentioned in authors work [74], α must be adjusted in accordance to the
measured profiles at the target location of the building. For the ground wall, surface roughness needs to be defined in accordance to the
terrain category for which the building is located. For example in ANSYS FLUENT [65], surface roughness at the ground is defined
based on sand grain height (ks ) and roughness length (zo ) relationship as given in Eq. (4) [94]. The parameter cs is known as the
roughness constant and takes a value between 0.5 and 1. The provided surface roughness and the resulting calculated sand grain height
ensures that the first cell height is larger than this value, as required in ANSYS Fluent [65].
9.793zo
ks = Eq. 5
cs
One of the key concerns of setting up the domain is the size definitions in terms of boundary distances from the target structure/
structures. Franke et al.’s [95] CFD guideline recommends using 5H (where H is the building height) distance from the inlet to the
target building and 15H for the downstream length for flow to develop fully. A distance of 5H is recommended on lateral sides. These
recommendations are also upheld by Tominaga et al. [92], who also mentions the lateral blockage ratio must be less than 3%, which is
often the target in wind tunnel experimental. A more recent study by Abu-Zidan et al. [96] showcased that domain configurations
presented in available guidelines [92,95] were overly conservative, which results in domain sizes with large cell counts. Using nu­
merical simulation evidence with the employment of RANS and LES models, Abu-Zidan et al. [96] showcased the importance of several
aspects which were not limited to frontal blockage and specified that domain size is problem specific and presented a framework for
optimisation of the domain. This optimisation of the domain was based on: (1) wind blocking effects due to short upstream length, (2)
flow recirculation that may occur due to a short downstream length, (3) global venturi effects on the cross-plane due to large blockage
ratios and (4) local venturi effects due to small blockage ratios, i.e. insufficient space between lateral and top surface of the domain.

Fig. 6. Mesh configuration setup for buildings where (a) building surface mesh and (b) inflation layers on surface and first cell height and (c) overal mesh refine­
ment plan.

8
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

Generally, the rule of thumb approach for tall building applications is to maintain blockage ratios to within 3% with a maximum of 5%
[76].
The computational grid is perhaps the most crucial aspect of a CFD simulation. The defined grid must be fine enough to resolve all
scales of interest for the particular problem at hand but also not too fine, as it could severely affect computational time. Based on the
authors experience in CFD modelling for buildings, four important mesh considerations need to be accounted for; (1) Surface cell count
for pressure distribution refinedness, (2) Inflation layers, their growth rate and first cell height in resolving wall-bounded flows, (3)
upstream mesh refinement for maintaining wind profile propagation and synthetic turbulence, (4) downstream refinement to capture
the formation of vortices and recirculation of flow. Fig. 6 displays a plane cut through the middle of a CFD domain where an isolated
super-tall structure is concerned. This particular mesh has been employed in the author’s previous work [74,97], where Fig. 6(a)
displays the surface mesh. The structure has circular cross-sections where the surface mesh was defined such that 80 cell divisions are
provided on the circumference. In contrast, Franke et al. [95] defined that a minimum of 10 cells must be provided per building face,
and it is clear that this is aimed at structures with flat surfaces. Fig. 6(b) displays the inflation layer where 20 prism cells were provided
with an average y+ avg < 0.35 and a maximum ymax was 0.81 [97]. In comparison, both guides Ref’s [92,95] state the importance of using
+

prism layers in unstructured mesh at walls but don’t provide strict guidelines other than stating the stretching ratio shouldn’t exceed
1.3 [95]. In the work presented in Ref. [97], the authors took a structured approach to determining the boundary layer height and first
cell thickness, which is elaborated in Table 1. A combination of theoretical calculations and RANS simulations are used for optimi­
sation. Fig. 6(c) displays the overall mesh where the domain is split into three regions. Region 1 is the closest to the surrounding
building and is refined so detached scales and vortices are appropriately captured. Region 2 extends from the inlet to the building and
further downstream. This refinement is mainly adopted to maintain approaching synthetic turbulence and preserve the wind profile.
Also, the downstream is refined to capture the larger scales of flow vortices. Finally, Region 3 inhibits medium-coarse meshing at the
volume further away from the region of interest and serves only to provide continuity of flow within the domain. This strategic mesh
setup is highly beneficial as it allows the user to concentrate most of the refinement at the region of interest. Using structured grids is
almost impossible for practical applications of CFD in tall building design. Most tall buildings have complex architectural features and
it is extremely challenging to provide structured grids where geometrical cross-sections change along its height. The use of unstruc­
tured grid arrangement is best adopted in modern applications.
Another important consideration in terms of grid generation is the cell type. In earlier implementations of CFD, hexahedral and
tetrahedral cells are often found. Franke et al. [95] recommend the prevalent use of hexahedral cells due to their ability to achieve
faster convergence and smaller truncation errors. However, the exclusive use of hexahedral meshes in a CFD model is often impossible
to implement, especially for structures of unorthodox geometries, as encountered in tall building designs. The use of hexahedral cells is
often combined with tetrahedral, where the latter is used in the region closer to the vicinity of the building. These grid arrangements
are often termed “hexahedral dominant” grids, as demonstrated in Ref’s [50,72,77]. Recently, the use of Polyhedral cells has gained

Table 1
Step-by-step guide to developing surface mesh for LES type simulations.

Steps Description

1. Determine first cell height The first cell height for LES type simulations needs to be stringent such that y+ ≤ 1. As a first estimation, the first cell
height is calculated via the Reynolds number followed by the wall shear stress. The steps are articulated below;
(i) Reynolds number, Re = ρUstream Dmin /μ, where ρ is the density of the fluid (air), Ustream is the maximum
stream velocity, usually encountered at the apex of the building, Dmin is the minimum cross-sectional
diameter and μ is the kinetic viscosity of fluid.
2
(ii) The wall shear stress τwall = Cf ρUstream /2 where Cf is the skin friction coefficient and is usually
0.2
approximated using Cf = 0.058 × Re− for external flows.
√̅̅̅̅̅̅̅̅̅
τwall
(iii) The friction velocity u∗ is calculated and given as u∗ = .
ρ
(iv) Then final the wall distance is calculated as y = y μ/u∗ ρ.
+

For example in the author’s publication [97] the apex velocity was 12 m/s and the smallest diameter (at the neck) was
0.08 m. The desired y+ ≤ 1 and following steps (i) to (iv) yielded a first cell height of roughly 2 × 10− 5 m.
2. Determine boundary layer Turbulent flow over a flat plate is used as an initial estimation for boundary layer thickness (δ) where it is calculated δ =
thickness 0.37 × Dmin /Re0.2
3. Number of Prism layers and The number of prism layers used needs to be decided. A low number of prism layers would result in a large growth rate in
growth rate order to cover the boundary layer. Franke et al. [95] recommends that the growth rate should not exceed 1.3. A large
number of prism layers could drastically increase the mesh count. Thus, an optimum number of cells need to be chosen and
this is usually done by constraining the growth rate and the number of cells required to cover the boundary layer. In the
author’s opinion, the growth rate should be kept to a minimum (ideally not exceeding 1.2) and maximum prism cells kept
to a maximum of 20. Once the boundary first cell layer height (y) and boundary layer thickness (δ) is known the growth
rate and number of prism layers can be calculated on a spreadsheet where the user has the ability to change the two
variables until the desired outcome is obtained.
4. Trial numerical simulation and The calculations provided in steps 1–3 are a good starting point to begin constructing the mesh. To consolidate the selected
readjustments cell size, it is recommended to run a RANS steady state simulation where the y+ can be directly obtained on the building
surface. This initial steady state RANS simulation should give an indication to the user if the provided cell quantities are
sufficient.
5. Final LES simulation When performing the final LES simulation it is important that y+ statistics are also monitored to ensure that the user
desired quantities are met. Also, the final RANS simulation which is used for confirmation can be used as the initial starting
solution for LES simulation ensuring good convergence for the simulation.

9
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

traction due to their capability to predict recirculating flows at a lower cell count and ability to achieve fast convergence of residuals
compared to tetrahedral cells [98]. Polyhedral cells have been successfully implemented across multiple applications of tall buildings,
such as in interference effects [99], aerodynamic shape modifications [85] and wind loads on building surfaces [74]. A comparison of
performance between Polyhedral cells and Hex-Tet grid for a super-tall building was compared in the authors work [100]. The mean
pressure around the circumference of the tall building was compared with experimental outcomes. These pressures were recorded at
two different height locations as presented in Fig. 7. As observed in Fig. 7, the Polyhedral mesh provided the closest predictions to the
experimental outcomes and only consists 1.7 million cells. In comparison the Hex-Tet mesh at even higher cell count (3.2 million cells),
couldn’t predict the negative pressure on the surface of the building. It must be noted that the cell quantities such as number of
inflation layers and first cell height of the grids remained consistent. The only changed values between them were the mesh density for
Region-1, Region-2 and Region-3 (as marked in Fig. 6). The 3.2 million Hex-Tet grid was comparable to the Polyhedral 1.7 million grid
in terms of all mesh density settings and yet, took longer in terms of computational time and presented poorer results in pressure
predictions.

3. Applications of CFD in tall buildings and structures


Some of the earliest applications of CFD on tall buildings were mostly used to determine wind loads & forces on the buildings. Most
CFD codes were implemented for other engineering disciplines and were mainly outside the scope of Civil/Structural Engineering
applications. CFD applications and research have evolved from validation of isolated buildings to more sophisticated modelling,
including aerodynamic shape modification studies, to more complex Fluid-Structure Interaction (FSI) in structural design applications.
Hence this section aims to encapsulate the progress of CFD application pertaining to tall buildings/structures.

3.1. Tall building validation studies


Validation studies on tall buildings primarily focus on building surface pressures and force coefficients. The CFD simulations are
validated against wind tunnel experiments where some of the earlier publications mainly utilised the CAARC building due to the
availability of experimental results in literature. Surface pressures are often compared by mean (cpm ) and RMS of fluctuations (the
standard deviation) (cp,rms ) pressure coefficients. In most literature, the pressure around the CAARC building is recorded, compared and
presented at 2/3rd the height of the building. Fig. 8 compares numerical results from CFD studies from Ref’s [50,84,91,101,102],
where scale-resolving turbulence models are employed. Table 2 displays the nomenclature of the results presented in Fig. 8 (where
curves C1–C6 are tagged after the Ref’s [50,84,91,101,102]) and describes the differences between the tests, which will be critical in
explaining the observed differences on the curves.
( )
A comparison of the windward face 0 < Dx < 1.5 shows that for almost all simulations, the mean pressure (cpm ) is close to one
another, with a few minor differences (Fig. 8(a)). However, it is noted that for C3, the windward pressures are higher than predicted by
their wind tunnel experiments [84]. Comparing the RMS pressure (cp.rms ) (Fig. 8(b)) on the windward face, it can be seen that C1
presents the smallest on the windward face with values of zero. The cp.rms on the windward face is an indication of the turbulence of the
approaching inflow profile, and since there is no turbulence for the smooth inflow, C1 presents a zero value. For cp.rms , the values
increase in ascending order from C4, C3, C6 and C5 as observed Fig. 8(b). It can be seen that although C3 has a higher turbulence
profile (Fig. 8(c)) compared to all other tests, it produces smaller cp.rms on the windward face compared to C5 and C6. The turbulence
intensity curve for C3, in theory, must have a higher cp.rms on the windward face compared to all other test cases but is not reflected in
( )
Fig. 8(b). On the crosswind face 1.5 < Dx < 2.5 in general, negative cpm is observed due to separation that occurs from the leading
edge. The magnitude of the negative pressure is the lowest for C1 as the approach flow is smooth and carries no turbulence. For test
cases with approaching inflow turbulence (C2–C6), the side pressure accumulates larger negative cpm . C3 and C6 display similar
negative cpm values even though they have the highest and smallest inflow turbulence, respectively. C5 once again shows the most

Fig. 7. Comparison of Polyhedral mesh and Hex-Tet mesh of a super-tall structure [100].

10
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

Fig. 8. Comparison of surface pressure measurements (at 2/3rd height) from CFD simulations of the CAARC building extracted from Ref’s [50,84,91,101,102] (a)
Mean pressure coefficient (cpm ), (b) RMS pressure coefficient (cp,rms ) and (c) Turbulence Intensity plots.

Table 2
Nomenclature and details of compared tests on CAARC building.

Test Ref Inflow type Turbulence Intensity Rating & Description Turbulence measured
location

C1 Zhang et al. [101] Smooth Smooth inflow turbulence NA


C2 Dagnew & Bitsuamlak [91] Turbulent The study employed a medium intensity turbulence profile At the inlet
(< 20%)
C3 Elshaer et al. [84] Turbulent The study employed a high intensity turbulence profile (> 30%) At the inlet
C4 Huang et al. [50] Turbulent The study employed a high turbulence intensity profile (> 30%) At the inlet
C5 The authors (Wijesooriya et al.) Turbulent The study employed a high turbulence intensity profile (> 30%) At the building location
[102]
C6 Zhang et al. [101] Turbulent The study employed a high turbulence intensity profile (> 35%) At the inlet

significant negative cpm on the side face. This trend is also observed for the cp.rms readings taken for all test cases. By observing C1
(which has no turbulence) and comparing to other test cases, it is clear that the side cp.rms is affected mainly by the approach flow
( )
turbulence. For the leeward face 2.5 < Dx < 4.0 , negative cpm are observed but their values are smaller in comparison to the side face.
This is also reflected in the cp.rms plots where the side face has larger values and smaller values are observed in the leeward face.
Generally, the mean and RMS pressure around the building plot location presents expected distribution trends. However, some
inconsistencies were observed. For example, turbulence intensity plots of C3, C4, and C5 present almost identical curves but present
differences in the side and leeward faces. In particular, C5 shows the largest RMS pressure values compared to all other test cases. In
theory, C3 should present larger values due to its high turbulence but instead displays a similar trend to C6 in both mean and RMS
pressure plots. The reason for such observation could be due to the loss of turbulence as the flow approaches the building. As noted in
Table 2, all studies record characteristic inflow data at the inlet, and only C5 presents inflow characteristics captured in an empty CFD

11
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

domain at the building location [102]. For Ref [102], both mean and RMS pressures were compared with their unique wind tunnel test
showed good similitude correlation. Given that inflow turbulence plays a significant role in accurate predictions of RMS pressure (as
mentioned by Daniels et al. [103]), it is important that CFD simulation inlet conditions must be validated at the building location. This
indicates that for all CFD work, all inlet conditions such as velocity profile, turbulence, power spectral density (PSD) of flow, inflow
turbulence integral length scales, cross-correlations, etc., must be measured at the building location. The requirement to measure
inflow characteristics at the building location of an empty domain was also discussed in Section 2.2.
Buildings with irregular geometries have also been studied and validated for pressure measurements. The authors (Mohotti et al.)
[64] studied a super-tall structure with varying circular cross-sections. The study was conducted to evaluate the capability of RANS
CFD models to predict salient flow features encountered in circular cross-sections. The authors (Mohotti et al.) [64] showed that RANS
turbulence models could predict windward and crosswind pressures but were poor at predicting leeward pressure. The use of the DDES
model improved leeward pressure but was still not satisfactory. Further investigation in Ref. [74] revealed that for the same building in
Ref. [64], the use of LES modelling and polyhedral cells in the domain provided very accurate predictions compared to the experi­
mental outcomes. Yan & Li [104] also studied an elliptical-shaped slender structure where surface roughness and its effects on pressure
predictions were studied. Out of the employed scale-resolving turbulence models, LES was the most capable of predicting pressure,
drag and lift coefficients. Zheng et al. [105] conducted a study on high-rise rectangular building with and without balconies. It was
shown that RANS models could only accurately predict windward pressure coefficients, whilst an LES model was required for cases
where pressure was read at other faces and when wind attack angles were not normal to the wide face of the body. Finally, a wide array
of building shapes and types such as E − shape [106], L - shaped [107], and other arbitrary geometries [108] have been studied by
various researchers. These studies often discuss the pressure distributions of these unique shaped buildings including wake flow effects
on downstream structures.

3.2. Aerodynamic optimisation


Aerodynamic optimisation of structures entails the design of exterior building shapes and features to minimise wind-induced loads
in both alongwind and crosswind directions. Vibrations in modern buildings have become more problematic as modern construction
technology employs strong yet light materials to reduce the structure’s carbon footprint and overall construction cost. The main
concern in modern tall buildings is vortex-induced vibration (VIV), where the frequency of crosswind shedding of vortices coincides
with the structure’s natural frequency. Fig. 9(a) presents an idealisation of crosswind response for a structure with and without its
susceptibility to vortex-induced vibration. Fig. 9(b) presents real wind tunnel measurements for crosswind response of a 1000 m mega-
tall building [19]. Irwin [109] mentioned that an approach to avoiding cross-wind response is to stiffen the building so that the
location at which peak resonance from VIV occurs at much higher wind speeds which would be unlikely to be encountered in its design
lifetime. However, this would not be financially feasible; hence, the best approach is to mitigate and minimise vortex shedding
altogether by shape optimisation.
CFD studies on shape optimisation have been a popular area of research due to the fast modelling techniques at disposal. Some of
the earlier studies utilised 2D RANS modelling. Bernardini et al. [110] presented one of the earliest studies with a Kriging-based
approach citing that such a methodology is far more efficient than performing random experimental models. As illustrated in
Fig. 10(a), the baseline model and the control points (and their limits) are presented, providing the surrogate model to make the
modifications. Fig. 10(b) shows all the possible edge modifications within the user-defined criteria. Bernardini et al. [110] used the
Pareto front framework to update the Kriging models, thus only requiring 0.75% of the planned CFD models to reach an optimum
shape based on minimising drag and lift coefficient. Ding & Kareem [111] further enhanced the work in Ref. [110], where a co-Kriging
approach was utilised where both LES and RANS modelling was used to evaluate an optimum cross-sectional shape. Elshaer et al. [85]
used an Artificial Neural Network (ANN) based surrogate model trained based on 2D LES flow for aerodynamic optimisation. Once the
optimal aerodynamic shape was identified, a full 3D LES model was performed to verify the outcomes in an ABL flow. Using a square
cross-section as a base model, it was shown that the optimal solutions for drag and lift were lowered by 30% and 24% respectively.

Fig. 9. Vortex-induced vibration: (a) idealisation and (b) experimental measurements from Zheng et al. [19].

12
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

Fig. 10. Surrogate Kriging-based approach: (a) base geometry with variable points and connections for geometry modifications and (b) all possible geometry
modifications for optimisation [110].

In a more active approach, Alminhana et al. [112] investigated the effects of corner modifications on the CAARC building where
both experimental and numerical (3D LES simulations) were conducted. Generally, recess and chamfered corners can drastically
reduce both drag and RMS lift coefficients exerted on the building [112]. Agrawal et al. [113] performed tests on multiple buildings
with shape modifications, set-backs and tapered structures, which closely follow the experimental work of Tanaka et al. [114]. It was
shown using a high-fidelity RANS turbulence model (SSTkω) surface pressure, drag and base moments were evaluated with some
discrepancies compared to the experimental outcomes. However, the simulation results were achieved in a computational
time-efficient manner due to the use of RANS models. Still, Agrawal et al. [113] stressed that such an approach should only be used as a
preliminary tool especially due to its incapability to acquire crosswind quantities.
A thorough review performed by Sharma et al. [115] covers important considerations on aerodynamic shape modifications and
their effects on structures. Sharma et al. [115] also performed independent CFD analysis to further validate and showcase the
effectiveness of corner-modified, tapered and set-backed building designs. It must be concluded that, from a structural engineer’s
perspective, shape optimisation is usually achieved to limit drag and RMS lift forces for robust structural design. The Structural en­
gineer and Wind engineer needs to work closely with the architect to ensure that an optimal design is achieved. However, serviceability
limit state (SLS) design such building comfort in terms of occupancy comfort is overlooked but is the most critical design criteria. This
is usually limited by maximum deflection but more importantly, maximum allowable accelerations the structure sustains in SLS
conditions [116]. Thus, aerodynamic optimisation has a key role to play in the structure’s design as that outlined in most Irwin [2].
Creating openings in the building is a common strategy designers often employ to reduce wind-induced vibrations and forces
exerted on the structure [117]. LES simulations with the help of ANN, as outlined by Elshaer & Bitsuamlak [118], can be used to
optimise opening locations (Fig. 11(a)) on a building to reduce overall wind-induced forces of up to 47%. One of the drawbacks of
conventional opening methods is that available floor space and area are significantly reduced thus reducing useable area (Fig. 11(a)).
New building designs, such as the 432 Park Avenue building, in New York, employ building openings but at service floors. The building
consists of five openings throughout its height and these are achieved by removing the façade elements at these floors without altering
the structural design. The effectiveness of such an approach (Fig. 11(b)) was investigated by a recent numerical CFD study by Marsland
et al. [119], where 18 different configurations of openings are simulated where (a) single layer openings, (b) double layer openings and
(c) multi-layer openings are concerned. It was shown that where only a single-layer opening is concerned, the most optimum location is

Fig. 11. Opening configurations (a) conventional building openings studied in Ref. [118] and (b) Service floor opening with structural elements intact [119].

13
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

at 70% of the building height, and optimal configurations for multilayer openings are discussed. These optimisations were based on
ISO standards’ acceleration and deflection comfort criteria, as outlined in Marsland et al. [119]. More importantly, it was discussed
that reductions in crosswind induced vibrations through openings were mainly achieved due to the disruption on synchronised vortex
shedding. This is expected as at the levels where façade elements don’t exist, the cross-sectional shape changes and thus, technically,
acts as a geometric modification (see Fig. 11(b)). In the alongwind direction, forces are reduced simply due to the decrease in projected
area and free passage of flow through the building.

3.3. Shielding effects and city modelling


Early validation and fundamental studies of wind flow behaviour around structures have meant that most available literature
focuses on isolated building cases. However, it’s understood that super-tall buildings are usually found in dense metropolitan areas.
Thus, surrounding buildings and their influence must also be accounted for. Codes of practice such as the ASCE/SEI 49-21 [120]
mentions the use of “proximity models”, which include surrounding buildings or terrain that may adversely affect the flow of the
approach wind on the subject structure. Proximity models can be made of foam and do not need full architectural features as long as
they are rigid [120]. In general, as mentioned in the code [120], the extent of proximity models is difficult to predict and is based on a
case-by-case scenario and evaluated by the wind engineer. Usually, significant buildings within a radius of 300 m–1000 m (the
variation depends on the size of the tunnel) to the subject building is implemented in the tunnel [121].
Dagnew and Bitsuamlak [91] presented a study of upwind building interference where an isolated building (case 1) is compared
with an interfering building of half the height (case 2) and an interference building of the same height of the subject structure (case 3).
The numerical CFD LES model was validated against the experimental outcomes, and the visual outputs were used to study the
complex interference flow phenomenon. All force coefficients were observed to have decreased for the half-height building (case 2)
compared to an isolated building case. However, for case 3, the subject structure experienced larger RMS crosswind forces. This was
explained through CFD observations, where it was described that flow remained separated (from vortex shedding) of the interference
building due to the proximity of the subject structure. This effect consequently resulted in a wider wake in the subject structure, as
reported by Dagnew and Bitsuamlak [91]. Observing the average velocity plots (Fig. 12), it’s clear that for case 3, separation at the
upstream building and recirculation bubble reattachment occurs at the subject building, which may indicate the higher RMS crosswind
forces. For case 2, the same is not observed as the cross-sectional plane presented (Fig. 12(b)) is at H/3, and the interference building
height is H/2. At a cross-sectional plane close to the apex, flow separates not only from the sides but also at the top thus leading to
smaller side vortex shedding. The authors speculate that a cross-sectional plane lower than H/3 would show similarities to that
observed in Fig. 12(c).
Sharma et al. [122] also performed a study on the interference effects of two tall buildings (breadth, width and height ratio of 1:1:7)
with varying separation distances (S = X/B) of 1.5–10. It was shown that for a close separation distance (S = 1.5) the interfered
building experienced negative windward pressures. This was similar to what was observed in Dagnew and Bitsuamlak [91], who also
experienced negative windward pressures on the interfered building (with S = 0.67). Sharma et al. [122] showcased that when
buildings are close to one another in tandem, the two structures virtually behave as a singular large body. Due to the longer sides, the
negative pressure on the leeward face of the interfered building is generally lower than that of the isolated building. In general, it is
shown that as separation increases, and the interfered building moves out of the wake of the interference building, the drag forces
increase until the two are so far apart that there are negligent effects of shielding (S = 12 ∼ 16) [122]. Similar observations were also
made by Shirzadeh et al. [123], who studied interference effects such that the interfering building is situated in tandem, staggered with
a combination of either upstream or downstream.
Most shielding effect studies are concerned with only a single interference building. However, Elshaer et al. [99] presented a study
where they first examined wind effects on a tall building subjected to generic surrounding building obstruction configurations using
LES CFD. The generic surrounding configurations are shown in Fig. 13 (C0 to C3), along with their impact on the subject structure
concerning mean and RMS pressure coefficients. As observed from the mean pressure coefficients, the shielding effect on the interfered
building increases with the surrounding structure height. The concentrated positive wind pressure shifts further up, and the pressure in

Fig. 12. Cross-sectional plane average velocity contour plots at H/3 for (a) isolated building, (b) interference building of height H/2 and (c) interference building of
height H [91].

14
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

the shielded height becomes less positive, as previously observed and discussed by Ref’s [91,122,123]. The RMS pressure coefficient,
on the other hand, increases with shielding height and thus may induce higher localised maximum pressure on cladding faces.
However, as noted by Elshaer et al. [99], alongwind and crosswind base moments were reduced on the interfered structure with
increasing height of surrounding structures leading to lower accumulated forces on the subject structure. In essence, it can be said that
the existence of surrounding structures, generally, can lower the overall building forces as opposed to an isolated building case.
Perhaps the most intriguing part of Elshaer et al.’s [99] study is the discussion on the induced pressure of buildings due to the
development of Downtown Toronto from 1980 to 2015. Whilst, most buildings did experience an overall reduction in mean pressures,
some buildings accumulated slight increases in mean pressure coefficients due to new developments. This means that as a city evolves,
unfavourable effects can occur, especially if the building in question and its façade elements are old and needs servicing.
Although buildings may or may not be subjected to adverse effects, an evolving city can cause concern to pedestrian-level wind
comfort. The approach wind becomes more turbulent due to increased roughness. This turbulence, coupled with possible wind ac­
celeration due to tunnelling effects, can create issues for pedestrians, as presented in Ref’s [86,99]. The advantage of running CFD
models is the ability to visualise these flow patterns and to readily alter geometry and flow conditions to obtain an optimal solution.

3.4. Fluid-structure interaction


As presented in the preceding sections, a vast number of research has been conducted to validate and measure wind-induced
pressure loads on buildings using CFD. The ability to accurately replicate these wind-induced loads and flow mechanisms is a pre­
requisite to obtaining wind-induced structural responses of tall buildings. Fluid-Structure Interaction (FSI) simulations are when a
Multiphysics solver couples the fluid CFD solver with a structural finite element solver to obtain structural responses. FSI is the most
sophisticated numerical modelling technique employed to determine structural responses of super-tall buildings and is essentially the
numerical equivalence to experimental aeroelastic modelling. There are two types of FSI modelling: two-way FSI and one-way FSI, and
they differ by how the solvers are coupled and data is transferred between them. In a two-way FSI model, the interaction between the
fluid and its induced forces on a flexible body and the body’s response is shared between the two solvers within a time step in a time-
marching simulation. The wind-induced loads exerted on the body are transferred to a mechanical model, which then calculates the
deflection and transfers the data back to the fluid solver and this cycle repeats. In essence, the two-way FSI simulation is quite
comparable to MDOF aeroelastic modelling in experimental wind tunnel tests especially with the added benefit of measuring surface

Fig. 13. Comparison of mean and RMS pressure from shielding effect with building height changes for angle of attack of zero degrees [99].

15
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

pressures (which some experimental can also perform [7]), which is an inherent property of numerical CFD modelling. In a one-way
FSI simulation, the fluid solver transfers the data to the mechanical solver, where the equation of motion is solved, and deflections are
calculated, but no data is transferred back; hence denoted as a “one-way” simulation. This is equivalent to using a combination of HFBB
and HFPI in experimental modelling whereby structural analysis is performed once the wind induced loads are recorded to determine
the dynamic loads on the building (as explained in Section 1.1). Thus, in the one-way FSI technique the aeroelastic response and its
effects cannot be directly determined.
FSI modelling is extremely challenging in terms of computational power and implementation. As it can be perceived, two-way FSI
simulations present more realistic physical behaviour but are far more demanding than one-way FSI simulations. This is mainly due to
the constant feedback between the solvers within each time step. The difficulty of FSI modelling and its demanding nature means that
there is only a limited amount of literature available, especially in applying tall and super-tall buildings.
One of the earliest implementations of FSI simulations was presented by Braun & Awruch [71], where two-way FSI simulations
were conducted on the CAARC building. An LES model was employed with a smooth inflow profile (no turbulence). The partitioned
coupling scheme was based on their previous work, where the interested reader is directed to Ref. [124]. Braun & Awruch [71]
conducted six simulations where the response of the structure was captured at six reduced velocities (Vr ). The numerical results were
compared to the experimental results of Thepmongkorn et al. [125] which utilised a stick aero-elastic model to determine the building
response. It was shown that the FSI simulation was capable of accurately representing alongwind deflection quantities but poorly
estimated the cross-wind response. Braun & Awruch [71] recognised that the cross-wind response was highly sensitive to approaching
flow turbulence which they didn’t implement. Noticing the deficiencies of Braun & Awruch [71], Zhang et al. [101] presented an
uncoupled one-way FSI simulation study on the CAARC building, which employed synthetic inflow turbulence (DSRFG). Zhang et al.
[101] noted that two-way FSI simulations were computationally demanding and instead proposed a one-way FSI analysis which only
solved the fluid simulation once. This was achieved by altering the buildings’ natural frequency to obtain the structural response at
different reduced velocities. Reduced velocity (Vr ) is a function of apex wind speed (UH ), first natural frequency (f1 ) and a building
dimension (usually base width) (D) as that shown in Eq. (5).
UH
Vr = Eq. 6
f1 D
In experimental aeroelastic modelling the, f1 and D are constant and the tests are carried out for varying values of UH . Recorded
measurements such as deflection and acceleration are given as a function of UH which is then up-scaled to the prototype quantities
using scaling laws. Zhang et al. [101] noticed that Vr is also a function of f1 which can be easily altered in the structural domain. This
meant that the fluid flow is solved only once and the induced aerodynamic forces from a single CFD simulation is used repeatedly to
solve multiple structural simulations to determine the response at various Vr s. This was a far efficient method to that which was used in

Braun & Awruch [71] which performed new simulations at the measured Vr s. Furthermore, since the CFD solver was used only once,

Zhang et al. [101] employed a highly dense mesh consisting of 1.6 million tetra-hexahedral mesh, which was capable of resolving a
majority of the flow spectrum that is influential for the tall building flow. In comparison, Braun & Awruch [71] only employed 360 k
elements, which isn’t sufficient to resolve all the flow features. This was a necessary compromise that needed to be taken to obtain
two-FSI results within a feasible timeframe. Fig. 14 presents the alongwind mean, alongwind standard deviation and crosswind
standard deviation response of the CAARC building for the two-way FSI, one-way FSI and experimental. Whilst the alongwind
quantities of both numerical methods are comparable to the experimental, it is clear that the one-way FSI method is capable of
predicting the cross-wind responses of the building. This indicates that Zhang et al. [101] method of changing the structural natural
frequency is an applicable method. Furthermore, the two-way FSI method shows poor results on the cross-wind but this is due to the

Fig. 14. Comparison of CAARC building response between two-way FSI [71], uncoupled oneway FSI [101] and experimental [125] responses for: (a) alongwind mean,
(b) alongwind standard deviation and (c) crosswind standard deviation.

16
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

use of a smooth inflow as previously described. Although not discussed, Braun & Awruch’s [71] study may have demanded heavy
computational resources which is indicative from the mesh number used (360 k) and the number of data outputs obtained. In contrast,
Zhang et al. [101] mentioned the CFD simulation on 96 CPU cores took 122 h and 35 min per structural simulation on 12 CPU cores.
This is fairly time efficient given that superior numerical accuracy was obtained to which it was comparable to experimental outcomes.
Further attempts of employing two-way FSI simulations can be found in the works of Hasama et al. [126] and Feng et al. [127].
Hasama et al. [126] studied the structural response of a rectangular prism where the structure was simplified to a spring-mass system
(Fig. 15(a)). This spring-mass system is ideal for structures that inhibit complex structural elements in the prototype building. It was
shown that the proposed method was capable of estimating alongwind and crosswind responses, to a good degree accuracy, compared
to the experimental outcomes presented by Marukawa [128]. However, Hasama et al. [126] noted some discrepancies in the torsional
responses and concluded that the structural simplification could have been attributed towards these differences. It is known that spring
and mass models cannot replicate complex modes of vibrations to an acceptable level of accuracy. Zheng et al.’s [129] study showcased
the simplification of a tall building with a mass-spring system leading to discrepancies exceeding 16% for the fifth and sixth modes of
vibration. Furthermore, simplifying the structural model is counter-intuitive, given that one of the main advantages of employing a
numerical model to evaluate structural performance is the ability to represent the real structural model of the prototype building. Feng
et al. [127] performed two sets of studies where first a validation test case was presented for a 300 m tall building and then a 1000 m
tall building, for which only numerical outcomes were only discussed. Feng et al. [127] used the commercial code ANSYS [24] to
perform the simulations and two-way coupling. For the validation case, Feng et al. [127] only evaluated the structures’ performance at
three Vr s due to computational demand. Also, the LES model employed was coarse, where the average y+ was in the range of 15–70.

Furthermore, a simplified MDOF structural model was used, as shown in Fig. 15(b). As Feng et al. [127] discussed, the alongwind
measurements (and, to some extent, the crosswind) of standard deviation response showed disparities in results compared to the
experimental. Feng et al. [127] describe these as a combination of differences observed in the physical model and inflow wind
quantities. Whilst these explanations may hold true, the real difference between experimental and numerical outcomes is usually
attributed to poor mesh density. Poor mesh density leads to the inability to capture the necessary scales of flow required to describe the
accurate flow features around the structure. The study also further investigates the response of a 1000 m tall building where its
response with and without aerodynamic damping is studied for wind veering effects [127]. Péntek et al. [130] presented a
multi-partitioned, fully coupled FSI method where the ability to include the structural model plus additional devices such as mass
dampers as part of the structural system. The study showcases complex FSI simulations where multiple parametric studies on structural
model simplification and the addition of TMD’s are considered. Péntek et al. [130] used the CAARC building where the first two
fundamental sway modes of vibration are considered and the paper presents strictly numerical outcomes only.
The authors introduced a novel one-way FSI technique (Wijesooriya et al.) [131], where a numerical framework was introduced to
obtain wind-induced loads and responses in a time-efficient method. The framework consists of an ELES numerical modelling tech­
nique and presents a highly efficient pressure mapping technique (first introduced in the author’s previous study [102]) and was
tailored for civil, structural applications. The proposed method in Ref. [102] adopted a similar evaluation technique to Zhang et al.
[101]. The CFD model was run once, and the structural model was altered and run multiple times to achieve the required measurement
quantities at different Vr s. The difference was an efficient pressure mapping technique was proposed where the pressure load transfer

from CFD to Mechanical was performed via a MATLAB script. The pressure mapping from CFD to the mechanical solver was highly
optimised and fast. The Mechanical solver employed an implicit modal analysis which was also highly efficient and provided solutions
within minutes [102]. In its first implementation, the authors (Wijesooriya et al.) [102] performed tests on the CAARC building and
showcased its performance against experimental outcomes. Then it was validated for a super-tall structure of height 300 m (408 m with
stalk), as shown in Fig. 16(a), where an MDOF aeroelastic model was tested in the wind tunnel [97]. The one-way FSI technique
provided good numerical accuracy compared to the experimental outcome in both alongwind and crosswind responses. Also, the
numerical model was then used to optimise the building for acceleration and deflection control for occupant comfort, as discussed in

Fig. 15. Structural model idealisation (a) mass and spring system from Hasama et al. [126] and (b) simplified structural model from Feng et al. [127].

17
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

Ref. [97]. The authors (Wijesooriya et al.) [100] also presented a comparison of the one-way FSI method against a fully coupled
two-way FSI simulation similar to that which was adopted by Feng et al. [127] using ANSYS [24]. The structural model used for the
one-way and two-way FSI methods were identical and consisted of all structural elements, such as floor slabs, core walls, beams and
columns, as shown in Fig. 16(b). As discussed in Ref. [100], the CFD domain meshing and size were different. For the one-way FSI
simulation, a high-quality CFD mesh was provided with 1.7 million polyhedral cells. In contrast, the two-way FSI model had to employ
a hexahedral-tetrahedral mesh since a dynamic mesh was used (deformable and re-meshed where necessary) for the two-way FSI
simulation and consisted of 1.6 million elements. As the authors (Wijesooriya et al.) [100] pointed out, 1.6 million cells for the
hexahedral-tetrahedral mesh was insufficient to predict all flow features and required a minimum of 3.2 million mesh elements.
However, due to the constraint of computational time, this compromise had to be made much like that described in Feng et al. [127]. A
comparison of the tip deflection for alongwind and crosswind standard deviation is presented in Fig. 17. As observed, the one-way FSI
(u1w-FSI) curve is able to predict deflection quantities to a good degree of accuracy. The two-way FSI (u2w-FSI) has lower data points
and cannot predict important resonance features such as that observed in peaks labelled “1”, where vortex induced resonance occurs
for the building in the study [100]. To compare the computational demand, six two-way FSI analysis was performed where only six
data points were obtained for which a total of 4800 clock hours were required to perform in a HPC [100]. In comparison, the one-way
FSI simulation only took a total of 74 clock hours to obtain simulation results for eighteen different structural configurations.
It is evident that two-way FSI simulations are complex and require high computational power. Certain compromises need to be
made, such as mesh density, cell type, minimum cell y+ and even total flow time to achieve results within a reasonable time frame. This
means that even though “the in principal” superior two-way FSI simulation is supposed to yield better results, it fails to do so, and the
one-way FSI simulation leads to better outcomes. The only disadvantage the one-way FSI model inhibits is its inability to account for
negative aerodynamic damping, which usually occurs due to vortex induced resonance (VIR). However, as pointed out in the author’s
[100] reference, for most buildings prismatic shaped structures such as squares and rectangles, the point at which VIR occurs is when
velocities are in the range of 80 m/s to 100 m/s. This is beyond even ULS design speeds. The authors (Wijesooriya et al.) [100] further
dwell on the advantages of using a one-way FSI approach to which the interested reader is directed. In short, the compromise of not
including negative aerodynamic damping is warranted, especially if better flow features and the atmospheric boundary layer repli­
cation can be simulated, which is achieved in a one-way FSI simulation method. Ricci et al. [132] also presented a one-way FSI analysis
where the deflection of a tall building and the generated internal forces were evaluated and compared to wind tunnel experiments with
good correlation between each other.

4. Future of cfd modelling and other promising numerical techniques on super-tall structures
4.1. The Lattice-Boltzmann method (LBM)
Computational hardware has tremendously improved over the past decade, and access to high-performance computers is readily
available even to the average consumer. These advancements have generally meant that higher-order turbulence models such as LES
are more feasible and readily applied in engineering applications. The Lattice-Boltzmann method (LBM) has recently gained traction as
an alternative CFD numerical method. In LBM, the physics of the fluid flow is modelled using particle kinetic theory and has been found
to be a far more efficient modelling technique than the conventional CFD approach [133]. Applications of LBM-LES methods for wind
engineering are on the rise, where studies such as pedestrian-level wind of city models [134,135] and pollutant dispersion [136,137]
are just some examples. In the scope of tall buildings, Buffa et al. [138] presented an LBM-LES numerical study on a tall building where
both inflow characteristics and wind induced loads and pressures were validated. It was shown that these quantities were reproduced
with good numerical accuracy using the LBM-LES method. Han et al. [139] performed a study on a high rise building where the
LBM-LES method was compared with an experimental and a conventional CFD (i.e. Finite Volume – FVM-LES). Comparisons were
made for averaged time velocity profiles, turbulent kinetic energy profiles, flow structures & reattachment length and fluctuation

Fig. 16. Structural model for FSI simulation: (a) building height, (b) cut cross-section through stalk to showcase complex structural walls and slabs (c) first mode of
vibration and (d) third mode of vibration [100].

18
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

Fig. 17. Tip deflection comparison between experimental (Exp), one-way FSI (u1w-FSI) and two-way FSI (2w-FSI) for (a) standard deviation alongwind response and
(b) standard deviation crosswind response [100].

velocity characteristics. It was shown that LBM-LES was capable of providing good estimates to experimental outcomes with similar
accuracy to conventional FVM-LES modelling [139]. The study further dwells on the necessary mesh refinement requirements for the
LBM-LES to capture the required flow features. Santasmasas et al. [140] presented an LBM-LES simulation over a high-rise building,
the same structure as that which was compared by Han et al. [139]. Santasmasas et al. [140] adopted a zonal approach (much similar
to what is adopted in Embedded-LES simulations [74]) using LBM-LES in the region of interest. It was shown that the LBM-LES
provided good agreement with experimental outcomes for flow features of the structure. These outcomes were compared to a con­
ventional FVM-CFD simulation where a wall-modelled LES (WMLES) simulation was run. It was shown that the most demanding
LBM-LES simulation was still 25 times less expensive in comparison to the FVM-CFD approach [140].
In its current state, LBM is still mainly used as a research tool but has a great future potential to be used in practical applications of
Wind Engineering. Much like the FVM method a couple of decades ago, LBM CFD is still going through the validation phase, especially
in Wind Engineering applications. The main advantage of LBM is its faster computation time due to the use of collision theory and
particle kinematics which avoid direct solving of conservation equations as that encountered in traditional CFD code. It can also utilise
excellent parallel performance with modern computer hardware and scales well with CPUs and GPUs to perform their operations
[141]. LBM has been widely adopted on GPU architecture due to the parallelisation architecture available in modern hardware.
However, as pointed out in Ref. [140], one of the main drawbacks of LBM is the requirement to store large quantities of data for solved
quantities, sometimes drastically affecting the performance of large simulations. This was one of the main motivations for imple­
menting Embedded-LES using LBM in Santasmasas et al. [140]. LBM is still an evolving method and will gain maturity like FVM-CFD
and be applicable to a broader array of flow problems. For a theoretical background of the LBM, the interested reader is directed to
Krüger et al. [142].

4.2. Artificial intelligence (AI) approaches


Applications of Artificial Intelligence (AI) driven approaches in wind engineering have gained traction in the recent past. This is
mainly due to the availability of large data sets accumulated by years of experimental testing. AI techniques have been applied in
various engineering fields, such as structural health monitoring, systems identifications, design and optimisation, which are articu­
lately reviewed by Salehi & Burgueño [143]. Early studies of AI in wind engineering can be found, such as that of Martínez-Vázquez &
Rodríguez-Cuevas [144], who studied wind field data reproduction via an Artificial Neural Network (ANN) approach. Wu & Kareem
[145] used the ANN approach to study the dynamic non-linearity’s of bridge decks. Raissi et al. [146] presented a deep learning (DL)
algorithm capable of reproducing important flow quantities of circular cylinders undergoing VIV, such as the lift and drag forces,
without needing training data. Furthermore, the DL model was able to accurately replicate velocity and pressure contours [146]. AI

19
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

driven approaches are still at their early stage, especially where buildings are concerned, and some of the new studies, including their
outcomes and capabilities, are elaborated below.
Hu & Kwok [147] employed an ML approach to determine the mean and RMS pressure distribution of circular buildings stating that
such an approach would be much more efficient than wind tunnel and CFD techniques. They collected data from 47 studies and used
Decision Tree Regressor (DTR), Random Forest (RF), and Gradient Boosting Regression Tree (GBRT) ML algorithms where predictions
were compared with each other. Building surface pressures are affected by inflow turbulence and velocity, especially for circular
cross-sections. Thus, the ML algorithms were trained to predict quantities of pressure based on the Reynolds number and turbulence.
Hu & Kwok [147] showed that the ML algorithms could provide these estimates with reasonable accuracy, especially the GBRT al­
gorithm, as shown in Fig. 18. Lin et al. [148] also employed a ML based approach to determine cross-wind vibration displacements of
rectangular cylinders. From the ML algorithms compared, the GBRT algorithm provided the best estimates [148].
Kim et al. [149] presented an unsupervised ML algorithm to study pressure patterns on tall buildings where a clustering algorithm
was employed. The clustering algorithm was used to determine the importance of the observed pressure patterns and further un­
derstand the aerodynamic characteristics of the building. ML approaches on interference effects [150], wind flow patterns on tall
buildings [151], wind surface pressure measurements [152] and crosswind load effects [153], have been shown to provide results with
reasonable accuracy to experimental measurements.
From the literature above, it is clear that AI-driven approaches have a major role in the future of wind engineering. As shown, AI
algorithms are capable of predictive measurements and can be used to critically evaluate large datasets and understand underlying
patterns and causations of observed/acquired data. Although AI driven methods aren’t in the same class as CFD-based numerical
modelling, it is still a numerical approach capable of providing qualitative outputs. The main advantage of AI driven approach is its
ability to deliver results at a low and feasible cost, especially in comparison to wind tunnel methods. Furthermore, AI generated
numerical results are also much faster in comparison to CFD-based numerical modelling. Finally, the reliability of AI driven outputs
will only further improve as further data is collected and will be an excellent tool to complement existing methods such as wind tunnel
experiments and CFD-based numerical modelling of tall buildings.

5. Concluding remarks
This manuscript presents in-depth, comprehensive literature and critical discussion on the progress of Computational Fluid Dy­
namics and its applications on wind-sensitive tall structures. It is clear that over the past two decades, CFD and its application in tall
building design have progressed from a research tool to a more practical design tool that can be adopted with confidence. This has been
achieved through the valuable contributions of all researchers who have worked on the respective topics. First, most of the studies that
adopted CFD mainly focused on square/rectangular cylinders or simplistic geometries classified as tall buildings. These early vali­
dation studies primarily focused on adopting various turbulence models and their applicability. In general, it was shown that RANS
turbulence models could predict averaged pressure quantities of the simplistic geometries but showed poor performance in RMS
pressure predictions. Furthermore, for geometries consisting of curved geometrical features or complex shapes, RANS turbulence
models could not provide good numerical estimates. These initial studies were mainly focused on RANS turbulence models due to the
lower computational demand. Scale-resolving turbulence models were necessary if accurate pressure predictions were required, but
this was achieved at a cost.
Turbulence modelling is at the heart of all CFD simulations; however, other important modelling challenges, such as generation of
inflow quantities were also critical. Accurately representing the atmospheric boundary layer flow in the numerical domain is a pre­
requisite, which is also evident in experimental testing. Various techniques were discussed within the manuscript, but a significant
emphasis was placed on synthetic turbulence generation methods due to their popularity and low computational headroom. Synthetic
turbulence models which were explicitly made to recreate atmospheric boundary layer flow and its applications for building design
were made by various researchers such as the DSRFG [80], CDRFG [83] and NSRFG [90] methods (to name a few). These studies were
pivotal in the progression of CFD techniques on tall buildings as the atmospheric boundary layer representation in the CFD domain was
achieved. Another aspect of CFD modelling is attributed towards the domain creation and mesh generation techniques. The latter is
perhaps the more important of the two, as a high-quality grid needs to be capable of resolving all necessary scales of flow for the

Fig. 18. Surface pressure predictions based on Re and Ti for (a) mean pressure and (b) RMS Pressure [147]. (The markers are the experimental outcomes and solid line is
GBRT algorithm prediction).

20
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

problem at hand. Several grid generation techniques including cell types were discussed namely tetrahedral, hexahedral and poly­
hedral shapes. Recent literature suggest that polyhedral cells are able to provide good quality mesh (similar to that of hexahedral) but
with a much lower cell count [98]. Polyhedral are highly suitable for complex geometric structures where hexahedral meshing is
difficult to implement. The importance of refinement regions including local mesh sizes on the subject wall was discussed with the aid
of the author’s previous work [64,74]. The y+ criteria is important when implementing LES type models and the manuscript provides
a method by which an initial estimate can be achieved through calculations and some simplified RANS simulations. The effects of not
correctly capturing the wall boundary layer flow for curved surfaces were discussed and presented.
These modelling techniques and their applications through validation work performed by various researchers over the years were
then discussed based on four important topics. Early studies were focused on pressure validation studies, and it was shown that the
CAARC building was the most popular amongst researchers due to the availability of experimental results. Some numerical results of
various studies were cross-examined (for the CAARC building) and showcased the nuances and effects of how inflow quantities can
drastically affect pressure distribution, especially for RMS quantities. The importance of measuring inflow quantities in an empty
domain at the building location was highlighted. Structures of irregular geometries of differing cross-sectional shapes and tapered (or
setback) profiles were also studied, showcasing the ability of CFD to be successfully applied to complex-shaped structures. The dis­
cussion naturally led to shape optimisation strategies and implementation in CFD. Algorithm-based and artificial intelligence
frameworks for achieving optimal shapes were discussed, predominantly using 2D modelling techniques. Furthermore, studies on
shielding effects, city modelling and the effectiveness of using a CFD approach were discussed. The ability to measure and acquire data
without being limited to instrumentation and readily changing building and surrounding features meant that CFD was suitable for
optimisation studies.
A significant emphasis was made on Fluid-Structure Interaction (FSI) simulations in this manuscript, as hardly any other review
articles by other researchers [52–55] have covered this aspect in detail. The benefit of using a computational method of numerical
aeroelastic modelling is that the full structural model of the prototype model can be used. In experimental aeroelastic models, an
MDOF model of limited vibration modes is implemented. An experimental model may be hard to re-create for some structures where
complex vibrational modes (especially coupled) exist. Furthermore, creating such a model requires expertise and time. As Irwin et al.
[109] mentioned, a full aero-elastic model is usually reserved towards the end of the building design. In contrast, the full prototype
building is generally available numerically as it is often common practice to develop a full FE model for structural analysis. It was
discussed that there were two types of FSI modelling: coupled (two-way) or uncoupled (one-way) simulations. The former is a more
accurate representation in the physical sense, but harder to implement. The computational demand required for two-way FSI is such
that sacrifices in mesh quality and statistics must be implemented to obtain results within a feasible time frame. This naturally led to
poorer overall outcomes, and it was shown that most research uses one-way FSI techniques to provide accurate results to that of
aeroelastic experimental testing. The authors [97,100] proposed an innovative one-way FSI techniques, which seem to be the most
viable option in the current state.
Finally, the paper discusses future modelling prospects of CFD, such as the Lattice-Boltzmann Method (LBM) and other numerical
techniques in the form of Artificial Intelligence (AI). The section discussed a brief overview of the two LBM and AI and their potential
within the scope of wind engineering simulations for tall building design. Early studies and recent literature suggest both methods to be
extremely promising, with the capability to be fully adopted in practical applications in the near future. Clearly, numerical techniques
are here to stay and are ever-improving. These numerical techniques can be used to optimise designs further and support experimental
methods which have long stood as the primary design tool in wind engineering design of tall buildings.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

References
[1] S. Lamb, K.C.S. Kwok, D. Walton, Occupant comfort in wind-excited tall buildings: motion sickness, compensatory behaviours and complaint, J. Wind Eng. Ind.
Aerod. 119 (2013) 1–12.
[2] P. Irwin, Wind engineering challenges of the new generation of super-tall buildings, J. Wind Eng. Ind. Aerod. 97 (2009) 328–334.
[3] M. Jensen, The model-law for phenomena in natural wind, Ingenioren Int. Ed. 2 (1958) 121–128.
[4] L. Cochran, R. Derickson, A physical modeler’s view of Computational Wind Engineering, J. Wind Eng. Ind. Aerod. 99 (2011) 139–153.
[5] T. Tschanz, A.G. Davenport, The base balance technique for the determination of dynamic wind loads, J. Wind Eng. Ind. Aerod. 13 (1983) 429–439.
[6] J. Xie, P.A. Irwin, Application of the force balance technique to a building complex, J. Wind Eng. Ind. Aerod. 77 & 78 (1998) 579–590.
[7] C. Pozzuoli, G. Bartoli, U. Peil, M. Clobes, Serviceability wind risk assessment of tall buildings including aeroelastic effects, J. Wind Eng. Ind. Aerod. 123
(2013) 325–338.
[8] J.D. Holmes, Mode shape corrections for dynamic response to wind, Eng. Struct. 9 (3) (1987) 210–212.
[9] W. Boggs Daryl, A. Peterka Jon, Aerodynamic model tests of tall buildings, J. Eng. Mech. 115 (3) (1989) 618–635.
[10] M.F. Huang, K.T. Tse, C.M. Chan, K.C.S. Kwok, P.A. Hitchcock, W.J. Lou, G. Li, An integrated design technique of advanced linear-mode-shape method and
serviceability drift optimization for tall buildings with lateral–torsional modes, Eng. Struct. 32 (8) (2010) 2146–2156.
[11] K.T. Tse, P.A. Hitchcock, K.C.S. Kwok, Mode shape linearization for HFBB analysis of wind-excited complex tall buildings, Eng. Struct. 31 (3) (2009) 675–685.

21
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

[12] L. Zou, T. Shi, J. Song, C.S. Cai, Application of the high-frequency base balance technique to tall slender structures considering the effects of higher modes,
Eng. Struct. 151 (2017) 1–10.
[13] W. Cui, L. Caracoglia, Physics-based method for the removal of spurious resonant frequencies in high-frequency force balance tests, J. Struct. Eng. 142 (2)
(2016), 04015129.
[14] X. Chen, A. Kareem, Dynamic wind effects on buildings with 3D coupled modes: application of high frequency force balance measurements, J. Eng. Mech. 131
(11) (2005) 1115–1125.
[15] L. Rosa, G. Tomasini, A. Zasso, A.M. Aly, Wind-induced dynamics and loads in a prismatic slender building: a modal approach based on unsteady pressure
measurements, J. Wind Eng. Ind. Aerod. 107–108 (2012) 118–130.
[16] A.M. Aly, Pressure integration technique for predicting wind-induced response in high-rise buildings, Alex. Eng. J. 52 (4) (2013) 717–731.
[17] D. Huang, Z. Ledong, D. Quanshun, Z. Xue, C. Wei, Aeroelastic and aerodynamic interference effects on a high-rise building, J. Fluid Struct. 69 (2017)
355–381.
[18] L. Wang, S. Liang, G. Huang, J. Song, L. Zhang, Improved expression for across-wind aerodynamic damping ratios of super high-rise buildings, J. Wind Eng.
Ind. Aerod. 176 (2019) 263–272.
[19] C. Zheng, Z. Liu, T. Wu, H. Wang, Y. Wu, X. Shi, Experimental investigation of vortex-induced vibration of a thousand-meter-scale-mega-tall building, J. Fluid
Struct. 85 (2019) 94–109.
[20] F. Hou, P. Sarkar, Time-domain model for prediction of generalized 3DOF buffeting response of tall buildings using 2D aerodynamic sectional properties, Eng.
Struct. 232 (2021), 111847.
[21] G. Solari, Wind Loading of structures: framework, phenomena, Tools and codification, Structures 12 (2017) 265–285.
[22] Y. Zhou, A. Kareem, Aeroelastic balance, J. Eng. Mech. 129 (3) (2003) 283–292.
[23] L. Wang, S. Liang, G. Huang, J. Song, L. Zou, Investigation on the unstability of vortex induced resonance of high-rise buildings, J. Wind Eng. Ind. Aerod. 175
(2018) 17–31.
[24] Ansys®, Academic research fluent, Release 18 (2017).
[25] B. Blocken, Computational wind engineering: theory and applications, in: C.C. Baniotopoulos, C. Borri, T. Stathopoulos (Eds.), Environmental Wind
Engineering and Design of Wind Energy Structures, Springer Vienna, Vienna, 2011, pp. 55–93.
[26] B. Blocken, 50 years of computational wind engineering: past, present and future, J. Wind Eng. Ind. Aerod. 129 (2014) 69–102.
[27] B. Blocken, T. Stathopoulos, J.P.A.J. van Beeck, Pedestrian-level wind conditions around buildings: review of wind-tunnel and CFD techniques and their
accuracy for wind comfort assessment, Build. Environ. 100 (2016) 50–81.
[28] M. Maing, Superblock transformation in Seoul Megacity: effects of block densification on urban ventilation patterns, Landsc. Urban Plann. 222 (2022),
104401.
[29] P. Ding, X. Zhou, H. Wu, Q. Chen, An efficient numerical approach for simulating airflows around an isolated building, Build. Environ. 210 (2022), 108709.
[30] M. Alizadeh, S.M. Sadrameli, Numerical modeling and optimization of thermal comfort in building: central composite design and CFD simulation, Energy
Build. 164 (2018) 187–202.
[31] X. Shan, N. Luo, K. Sun, T. Hong, Y.-K. Lee, W.-Z. Lu, Coupling CFD and building energy modelling to optimize the operation of a large open office space for
occupant comfort, Sustain. Cities Soc. 60 (2020), 102257.
[32] Y. Du, B. Blocken, S. Pirker, A novel approach to simulate pollutant dispersion in the built environment: transport-based recurrence CFD, Build. Environ. 170
(2020), 106604.
[33] R. Zhao, S. Liu, J. Liu, N. Jiang, Q. Chen, Generalizability evaluation of k-ε models calibrated by using ensemble Kalman filtering for urban airflow and
airborne contaminant dispersion, Build. Environ. 212 (2022), 108823.
[34] Y. Tominaga, T. Stathopoulos, CFD simulation of near-field pollutant dispersion in the urban environment: a review of current modeling techniques, Atmos.
Environ. 79 (2013) 716–730.
[35] S. Murakami, Current status and future trends in computational wind engineering, J. Wind Eng. Ind. Aerod. 67–68 (1997) 3–34.
[36] W. Rodi, On the simulation of turbulent flow past bluff bodies, J. Wind Eng. Ind. Aerod. 46–47 (1993) 3–19.
[37] H. Lübcke, S. Schmidt, T. Rung, F. Thiele, Comparison of LES and RANS in bluff-body flows, J. Wind Eng. Ind. Aerod. 89 (14) (2001) 1471–1485.
[38] W. Rodi, Comparison of LES and RANS calculations of flow around bluff bodies, Wind Eng. Ind. Aerodyn. 69–71 (1997) 55–75.
[39] A.G. Kravchenko, P. Moin, Numerical studies of flow over a circular cylinder at ReD=3900, Phys. Fluids 12 (2) (2000) 403–417.
[40] P. Catalano, M. Wang, G. Iaccarino, P. Moin, Numerical simulation of the flow around a circular cylinder at high Reynolds numbers, Int. J. Heat Fluid Flow 24
(4) (2003) 463–469.
[41] D.C. Wilcox, Reassessment of the scale-determining equation for advanced turbulence models, AIAA J. 26 (1988) 1299–1310.
[42] F.R. Menter, Two-equation eddy-viscosity turbulence model for engineering applications, AIAA J. 32 (8) (1994) 1598–1605.
[43] P.R. Spalart, W.-H. Jou, M. Strelets, S.R. Allmaras, Comments on the feasibility of LES for wings, and on a hybris RANS/LES approach, in: Proceedings of the
First AFOSR International Conference on DNS/LES, 1997.
[44] P.R. Spalart, S. Deck, M.L. Shur, K.D. Squires, M.K. Sterlets, A. Travin, A new version of detached-eddy simulation, resistant to ambiguous grid densities, Theor.
Comput. Fluid Dynam. 26 (6) (2006) 523–550.
[45] M. Strelets, Detached Eddy Simulation of Massively Separated Flows. AIAA Paper 2001-0879, 2001.
[46] F.R. Menter, M. Kuntz, R. Langtry, Ten years of industrial experience with the SST turbulence model, in: K. Hanjalic, Y. Nagano, M. Tummers (Eds.),
Turbulence, Heat and Mass Transfer 4, Begell House Inc., Turkey, 2003.
[47] A. Mochida, S. Murakami, M. Shoji, Y. Ishida, Numerical simulation of flowfield around Texas Tech building by large eddy simulation, J. Wind Eng. Ind. Aerod.
46–47 (1993) 455–460.
[48] R. Panneer Selvam, Computation of flow around Texas Tech building using k-ε and Kato-Launder k-ε turbulence model, Eng. Struct. 18 (11) (1996) 856–860.
[49] R. Panneer Selvam, Computation of pressures on Texas Tech University building using large eddy simulation, J. Wind Eng. Ind. Aerod. 67–68 (1997) 647–657.
[50] S. Huang, Q.S. Li, S. Xu, Numerical evaluation of wind effects on a tall steel building by CFD, J. Constr. Steel Res. 63 (5) (2007) 612–627.
[51] R.L. Wardlaw, G.F. Moss, A Standard Tall Building Model for the Comparison of Simulated Winds in Wind Tunnels, vol. 25, 1970. CAARC, CC662. Tech.
[52] F. Hou, M. Jafari, Investigation approaches to quantify wind-induced load and response of tall buildings: a review, Sustain. Cities Soc. 62 (2020), 102376.
[53] M. Jafari, A. Alipour, Methodologies to mitigate wind-induced vibration of tall buildings: a state-of-the-art review, J. Build. Eng. 33 (2021), 101582.
[54] Y. Abu-zidan, P. Mendis, D. Mohotti, S. Fernando, Wind design of tall buildings: the state of the art, Electron. J. Struct. Eng. 22 (2022) 53–71.
[55] M.S. Thordal, J.C. Bennetsen, H.H.H. Koss, Review for practical application of CFD for the determination of wind load on high-rise buildings, J. Wind Eng. Ind.
Aerod. 186 (2019) 155–168.
[56] J. Blazek, Chapter 3 - principles of solution of the governing equations, in: Computational Fluid Dynamics: Principles and Applications, third ed., Butterworth-
Heinemann, Oxford, 2015, pp. 29–72.
[57] C. Diaz-Daniel, S. Laizet, J.C. Vassilicos, Direct numerical simulations of a wall-attached cube immersed in laminar and turbulent boundary layers, Int. J. Heat
Fluid Flow 68 (2017) 269–280.
[58] W. Chen, C. Ji, D. Xu, Z. Zhang, Three-dimensional direct numerical simulations of vortex-induced vibrations of a circular cylinder in proximity to a stationary
wall, Phys. Rev. Fluids 7 (4) (2022), 044607.
[59] W. Chen, C. Ji, D. Xu, M.M. Alam, Three-dimensional direct numerical simulations of two interfering side-by-side circular cylinders at intermediate spacing
ratios, Appl. Ocean Res. 123 (2022), 103162.
[60] M. Zhao, A.-A. Mamoon, H. Wu, Numerical study of the flow past two wall-mounted finite-length square cylinders in tandem arrangement, Phys. Fluids 33 (9)
(2021), 093603.
[61] M. Tsuchiya, S. Murakami, A. Mochida, K. Kondo, Y. Ishida, Development of a new k− ε model for flow and pressure fields around bluff body, J. Wind Eng. Ind.
Aerod. 67–68 (1997) 169–182.

22
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

[62] F.-Q. Meng, B.-J. He, J. Zhu, D.-X. Zhao, A. Darko, Z.-Q. Zhao, Sensitivity analysis of wind pressure coefficients on CAARC standard tall buildings in CFD
simulations, J. Build. Eng. 16 (2018) 146–158.
[63] Y. Tominaga, Flow around a high-rise building using steady and unsteady RANS CFD: effect of large-scale fluctuations on the velocity statistics, J. Wind Eng.
Ind. Aerod. 142 (2015) 93–103.
[64] D. Mohotti, K. Wijesooriya, D. Dias-da-Costa, Comparison of Reynolds Averaging Navier-Stokes (RANS) turbulent models in predicting wind pressure on tall
buildings, J. Build. Eng. 21 (2019) 1–17.
[65] Ansys Inc, ANSYS Fluent User’s Guide, ANSYS, Inc. Southpointe 275 Technology Drive, Canonsburg, PA, 2013.
[66] J. Smagorinsky, General circulation experiments with the primitive equations, Mon. Weather Rev. 91 (3) (1963) 99–164.
[67] M. Germano, U. Piomelli, P. Moin, W.H. Cabot, A dynamic subgrid-scale eddy viscosity model, Phys. Fluid. 3 (7) (1991) 1760–1765.
[68] D.K. Lilly, A proposed modification of the Germano subgrid-scale closure method, Phys. Fluid. 4 (3) (1991) 633–635.
[69] W.-W. Kim, S. Menon, Application of the localized dynamic subgrid-scale model to turbulent wall-bounded flows, in: 35th Aerospace Sciences Meeting,
NVAmerican Institute of Aeronautics and Astronautics, Reno, 1997. Technical Report AIAA-97-0210.
[70] S. Murakami, Comparison of various turbulence models applied to a bluff body, J. Wind Eng. Ind. Aerod. 46 & 47 (1993) 21–36.
[71] A.L. Braun, A.M. Awruch, Aerodynamic and aeroelastic analyses on the CAARC standard tall building model using numerical simulation, Comput. Struct. 87
(9–10) (2009) 564–581.
[72] B.W. Yan, Q.S. Li, Inflow turbulence generation methods with large eddy simulation for wind effects on tall buildings, Comput. Fluid 116 (2015) 158–175.
[73] S. Murakami, Overview of turbulence models applied in CWE—1997, J. Wind Eng. Ind. Aerod. 74–76 (1998) 1–24.
[74] K. Wijesooriya, D. Mohotti, K. Chauhan, D. Dias-da-Costa, Numerical investigation of scale resolved turbulence models (LES, ELES and DDES) in the assessment
of wind effects on supertall structures, J. Build. Eng. 25 (2019), 100842.
[75] E.D. Obasaju, Measurement of forces and base overturning moments on the CAARC tall building model in a simulated atmospheric boundary layer, J. Wind
Eng. Ind. Aerod. 40 (2) (1992) 103–126.
[76] J.D. Holmes, S.A. Bekele, Wind Loading of Structures, fourth ed., CRC Press, 2021.
[77] M.S. Thordal, J.C. Bennetsen, S. Capra, H.H.H. Koss, Engineering approach for a CFD inflow condition using the precursor database method, J. Wind Eng. Ind.
Aerod. 203 (2020), 104210.
[78] A. Smirnov, S. Shi, I. Celik, Random flow generation technique for large eddy simulations and particle-dynamics modeling, J. Fluid Eng. 123 (2) (2001)
359–371.
[79] R.H. Kraichnan, Diffusion by a random velocity field, Phys. Fluid. 13 (1) (1970) 22–31.
[80] S.H. Huang, Q.S. Li, J.R. Wu, A general inflow turbulence generator for large eddy simulation, J. Wind Eng. Ind. Aerod. 98 (10) (2010) 600–617.
[81] H.G. Castro, R.R. Paz, V.E. Sonzogni, Generation of turbulent inlet velocity conditions for lage eddy simulations, Mecánica Comput. 30 (2011) 2275–2288.
[82] Y. Kim, I.P. Castro, Z.-T. Xie, Divergence-free turbulence inflow conditions for large-eddy simulations with incompressible flow solvers, Comput. Fluid 84
(2013) 56–68.
[83] H. Aboshosha, A. Elshaer, G.T. Bitsuamlak, A. El Damatty, Consistent inflow turbulence generator for LES evaluation of wind-induced responses for tall
buildings, J. Wind Eng. Ind. Aerod. 142 (2015) 198–216.
[84] A. Elshaer, H. Aboshosha, G. Bitsuamlak, A. El Damatty, A. Dagnew, LES evaluation of wind-induced responses for an isolated and a surrounded tall building,
Eng. Struct. 115 (2016) 179–195.
[85] A. Elshaer, G. Bitsuamlak, A. El Damatty, Enhancing wind performance of tall buildings using corner aerodynamic optimization, Eng. Struct. 136 (2017)
133–148.
[86] K. Adamek, N. Vasan, A. Elshaer, E. English, G. Bitsuamlak, Pedestrian level wind assessment through city development: a study of the financial district in
Toronto, Sustain. Cities Soc. 35 (2017) 178–190.
[87] A.A. Aliabadi, N. Veriotes, G. Pedro, A Very Large-Eddy Simulation (VLES) model for the investigation of the neutral atmospheric boundary layer, J. Wind Eng.
Ind. Aerod. 183 (2018) 152–171.
[88] Q. Yang, T. Zhou, B. Yan, M. Liu, P. Van Phuc, Z. Shu, LES study of topographical effects of simplified 3D hills with different slopes on ABL flows considering
terrain exposure conditions, J. Wind Eng. Ind. Aerod. 210 (2021), 104513.
[89] L. Patruno, M. Ricci, On the generation of synthetic divergence-free homogeneous anisotropic turbulence, Comput. Methods Appl. Mech. Eng. 315 (2017)
396–417.
[90] Y. Yu, Y. Yang, Z. Xie, A new inflow turbulence generator for large eddy simulation evaluation of wind effects on a standard high-rise building, Build. Environ.
138 (2018) 300–313.
[91] A.K. Dagnew, G.T. Bitsuamlak, Computational evaluation of wind loads on a standard tall building using LES, Wind Struct. 18 (5) (2014) 567–598.
[92] Y. Tominaga, A. Mochida, R. Yoshie, H. Kataoka, T. Nozu, M. Yoshikawa, T. Shirasawa, AIJ guidelines for practical applications of CFD to pedestrian wind
environment around buildings, J. Wind Eng. Ind. Aerod. 96 (10–11) (2008) 1749–1761.
[93] B. Blocken, Computational Fluid Dynamics for urban physics: importance, scales, possibilities, limitations and ten tips and tricks towards accurate and reliable
simulations, Build. Environ. 91 (2015) 219–245.
[94] B. Blocken, T. Stathopoulos, J. Carmeliet, CFD simulation of the atmospheric boundary layer: wall function problems, Atmos. Environ. 41 (2) (2007) 238–252.
[95] J. Franke, C. Hirsch, A. Jensen, H. Krüs, M. Schatzmann, P. Westbury, S. Miles, J. Wisse, N. Wright, Recommendations on the use of CFD in wind engineering,
in: COST Action C14, Impact of Wind and Storm on City Life Built Environment, von Karman Institute, Sint-Genesius-Rode, Belgium, 2004.
[96] Y. Abu-Zidan, P. Mendis, T. Gunawardena, Optimising the computational domain size in CFD simulations of tall buildings, Heliyon 7 (4) (2021), e06723.
[97] K. Wijesooriya, D. Mohotti, A. Amin, K. Chauhan, Wind loads on a super-tall slender structure: a validation of an uncoupled fluid-structure interaction (FSI)
analysis, J. Build. Eng. 35 (2021), 102028.
[98] M. Peric, S. Ferguson, The Advantage of Polyhedral Meshes, 2005. Technical report CD Adapco.
[99] A. Elshaer, A. Gairola, K. Adamek, G. Bitsuamlak, Variations in wind load on tall buildings due to urban development, Sustain. Cities Soc. 34 (2017) 264–277.
[100] K. Wijesooriya, D. Mohotti, A. Amin, K. Chauhan, Comparison between an uncoupled one-way and two-way fluid structure interaction simulation on a super-
tall slender structure, Eng. Struct. 229 (2021), 111636.
[101] Y. Zhang, W.G. Habashi, R.A. Khurram, Predicting wind-induced vibrations of high-rise buildings using unsteady CFD and modal analysis, J. Wind Eng. Ind.
Aerod. 136 (2015) 165–179.
[102] K. Wijesooriya, D. Mohotti, A. Amin, K. Chauhan, An uncoupled fluid structure interaction method in the assessment of structural responses of tall buildings,
Structures 25 (2020) 448–462.
[103] S.J. Daniels, I.P. Castro, Z.-T. Xie, Peak loading and surface pressure fluctuations of a tall model building, J. Wind Eng. Ind. Aerod. 120 (2013) 19–28.
[104] B.W. Yan, Q.S. Li, Detached-eddy and large-eddy simulations of wind effects on a high-rise structure, Comput. Fluid 150 (2017) 74–83.
[105] X. Zheng, H. Montazeri, B. Blocken, CFD simulations of wind flow and mean surface pressure for buildings with balconies: comparison of RANS and LES, Build.
Environ. 173 (2020), 106747.
[106] B. Bhattacharyya, K. Dalui Sujit, Experimental and numerical study of wind-pressure distribution on irregular-plan-shaped building, J. Struct. Eng. 146 (7)
(2020), 04020137.
[107] W. Yuan, Z. Wang, H. Chen, K. Fan, Numerical analyses of aerodynamic characteristics of integrated L-shaped high-rise building, Adv. Eng. Software 114
(2017) 144–153.
[108] P. Sanyal, S.K. Dalui, Effects of internal angle between limbs of “Y” plan shaped tall building under wind load, J. Build. Eng. 33 (2021), 101843.
[109] P. Irwin, J. Kilpatrick, J. Robinson, A. Frisque, Wind and tall buildings: negatives and positives, Struct. Des. Tall Special Build. 17 (5) (2008) 915–928.
[110] E. Bernardini, S.M.J. Spence, D. Wei, A. Kareem, Aerodynamic shape optimization of civil structures: a CFD-enabled Kriging-based approach, J. Wind Eng. Ind.
Aerod. 144 (2015) 154–164.
[111] F. Ding, A. Kareem, A multi-fidelity shape optimization via surrogate modeling for civil structures, J. Wind Eng. Ind. Aerod. 178 (2018) 49–56.

23
K. Wijesooriya et al. Journal of Building Engineering 74 (2023) 106828

[112] G.W. Alminhana, A.L. Braun, A.M. Loredo-Souza, A numerical-experimental investigation on the aerodynamic performance of CAARC building models with
geometric modifications, J. Wind Eng. Ind. Aerod. 180 (2018) 34–48.
[113] S. Agrawal, J.K. Wong, J. Song, O. Mercan, P.J. Kushner, Assessment of the aerodynamic performance of unconventional building shapes using 3D steady
RANS with SST k-ω turbulence model, J. Wind Eng. Ind. Aerod. 225 (2022), 104988.
[114] H. Tanaka, Y. Tamura, K. Ohtake, M. Nakai, Y. Chul Kim, Experimental investigation of aerodynamic forces and wind pressures acting on tall buildings with
various unconventional configurations, J. Wind Eng. Ind. Aerod. 107 (Supplement C) (2012) 179–191.
[115] A. Sharma, H. Mittal, A. Gairola, Mitigation of wind load on tall buildings through aerodynamic modifications: review, J. Build. Eng. 18 (2018) 180–194.
[116] F.A. Johann, M.E.N. Carlos, F.L.S. Ricardo, Wind-induced motion on tall buildings: a comfort criteria overview, J. Wind Eng. Ind. Aerod. 142 (2015) 26–42.
[117] Y.-G. Li, M.-Y. Zhang, Y. Li, Q.-S. Li, S.-J. Liu, Experimental study on wind load characteristics of high-rise buildings with opening, Struct. Des. Tall Special
Build. 29 (9) (2020), e1734.
[118] A. Elshaer, G. Bitsuamlak, Multiobjective aerodynamic optimization of tall building openings for wind-induced load reduction, J. Struct. Eng. 144 (10) (2018),
04018198.
[119] L. Marsland, K. Nguyen, Y. Zhang, Y. Huang, Y. Abu-Zidan, T. Gunawardena, P. Mendis, Improving aerodynamic performance of tall buildings using façade
openings at service floors, J. Wind Eng. Ind. Aerod. 225 (2022), 104997.
[120] (Asce), A.S.o.E, Wind Tunnel Testing for Buildings and Other Structures, 2022. ASCE/SEI 49-21.
[121] D. Boggs, A. Lepage, Wind Tunnel Methods, 2006. SP-240-6.
[122] A. Sharma, H. Mittal, A. Gairola, Detached-eddy simulation of interference between buildings in tandem arrangement, J. Build. Eng. 21 (2019) 129–140.
[123] M. Shirzadeh Germi, H. Eimani Kalehsar, Numerical investigation of interference effects on the critical wind velocity of tall buildings, Structures 30 (2021)
239–252.
[124] A.L. Braun, A.M. Awruch, A partitioned model for fluid–structure interaction problems using hexahedral finite elements with one-point quadrature, Int. J.
Numer. Methods Eng. 79 (5) (2009) 505–549.
[125] S. Thepmongkorn, K.C.S. Kwok, N. Lakshmanan, A two-degree-of-freedom base hinged aeroelastic (BHA) model for response predictions, J. Wind Eng. Ind.
Aerod. 49 (1999) 121–140.
[126] T. Hasama, T. Saka, Y. Itoh, K. Kondo, M. Yamamoto, T. Tamura, M. Yokokawa, Evaluation of aerodynamic instability for building using fluid–structure
interaction analysis combined with multi-degree-of-freedom structure model and large-eddy simulation, J. Wind Eng. Ind. Aerod. 197 (2020), 104052.
[127] C. Feng, G. Fu, M. Gu, Numerical simulation of wind veering effects on aeroelastic responses of thousand-meter-scale super high-rise buildings, J. Build. Eng.
46 (2022), 103790.
[128] H. Marukawa, N. Kato, K. Fujii, Y. Tamura, Experimental evaluation of aerodynamic damping of tall buildings, J. Wind Eng. Ind. Aerod. 59 (1996) 177–190.
[129] C. Zheng, Y. Xie, M. Khan, Y. Wu, J. Liu, Wind-induced responses of tall buildings under combined aerodynamic control, Eng. Struct. 175 (2018) 86–100.
[130] M. Péntek, A. Winterstein, M. Vogl, P. Kupás, K.-U. Bletzinger, R. Wüchner, A multiply-partitioned methodology for fully-coupled computational wind-
structure interaction simulation considering the inclusion of arbitrary added mass dampers, J. Wind Eng. Ind. Aerod. 177 (2018) 117–135.
[131] K. Wijesooriya, An uncoupled fluid-structure interaction numerical framework to estimate wind induced loads on super-tall structures, in: Faculty of
Engineering, The University of Sydney, Sydney, 2021.
[132] M. Ricci, L. Patruno, S.d. Miranda, B. Blocken, Towards LES as a design tool: wind loads assessment on a high-rise building, J. Wind Eng. Ind. Aerod. 180
(2018) 1–18.
[133] D.D. Ganji, S.H.H. Kachapi, Chapter 6 - natural, mixed, and forced convection in nanofluid, in: D.D. Ganji, S.H.H. Kachapi (Eds.), Application of Nonlinear
Systems in Nanomechanics and Nanofluids, William Andrew Publishing, Oxford, 2015, pp. 205–269.
[134] N. Onodera, T. Aoki, T. Shimokawabe, H. Kobayashi, Large-scale LES wind simulation using lattice Boltzmann method for a 10 km× 10 km area in
metropolitan Tokyo, Tsubame ESJ 9 (2013) 2–8.
[135] J. Jacob, P. Sagaut, Wind comfort assessment by means of large eddy simulation with lattice Boltzmann method in full scale city area, Build. Environ. 139
(2018) 110–124.
[136] L. Merlier, J. Jacob, P. Sagaut, Lattice-Boltzmann large-eddy simulation of pollutant dispersion in complex urban environment with dense gas effect: model
evaluation and flow analysis, Build. Environ. 148 (2019) 634–652.
[137] M.-F. King, A. Khan, N. Delbosc, H.L. Gough, C. Halios, J.F. Barlow, C.J. Noakes, Modelling urban airflow and natural ventilation using a GPU-based lattice-
Boltzmann method, Build. Environ. 125 (2017) 273–284.
[138] E. Buffa, J. Jacob, P. Sagaut, Lattice-Boltzmann-based large-eddy simulation of high-rise building aerodynamics with inlet turbulence reconstruction, J. Wind
Eng. Ind. Aerod. 212 (2021), 104560.
[139] M. Han, R. Ooka, H. Kikumoto, Validation of lattice Boltzmann method-based large-eddy simulation applied to wind flow around single 1:1:2 building model,
J. Wind Eng. Ind. Aerod. 206 (2020), 104277.
[140] M.C. Santasmasas, X. Zhang, B. Parslew, G.F. Lane-Serff, J. Millar, A. Revell, Comparison of lattice Boltzmann and Navier-Stokes for zonal turbulence
simulation of urban wind flows, Fluid 7 (2022), https://doi.org/10.3390/fluids7060181.
[141] S. Lenz, M. Schönherr, M. Geier, M. Krafczyk, A. Pasquali, A. Christen, M. Giometto, Towards real-time simulation of turbulent air flow over a resolved urban
canopy using the cumulant lattice Boltzmann method on a GPGPU, J. Wind Eng. Ind. Aerod. 189 (2019) 151–162.
[142] T. Krüger, H. Kusumaatmaja, A. Kuzmin, O. Shardt, G. Silva, E.M. Viggen, Numerical methods for fluids, in: T. Krüger, et al. (Eds.), The Lattice Boltzmann
Method: Principles and Practice, Springer International Publishing, Cham, 2017, pp. 31–58.
[143] H. Salehi, R. Burgueño, Emerging artificial intelligence methods in structural engineering, Eng. Struct. 171 (2018) 170–189.
[144] P. Martínez-Vázquez, N. Rodríguez-Cuevas, Wind field reproduction using neural networks and conditional simulation, Eng. Struct. 29 (7) (2007) 1442–1449.
[145] T. Wu, A. Kareem, Modeling hysteretic nonlinear behavior of bridge aerodynamics via cellular automata nested neural network, J. Wind Eng. Ind. Aerod. 99
(4) (2011) 378–388.
[146] M. Raissi, Z. Wang, M.S. Triantafyllou, G.E. Karniadakis, Deep learning of vortex-induced vibrations, J. Fluid Mech. 861 (2018) 119–137.
[147] G. Hu, K.C.S. Kwok, Predicting wind pressures around circular cylinders using machine learning techniques, J. Wind Eng. Ind. Aerod. 198 (2020), 104099.
[148] P. Lin, G. Hu, C. Li, L. Li, Y. Xiao, K.T. Tse, K.C.S. Kwok, Machine learning-based prediction of crosswind vibrations of rectangular cylinders, J. Wind Eng. Ind.
Aerod. 211 (2021), 104549.
[149] B. Kim, N. Yuvaraj, K.T. Tse, D.-E. Lee, G. Hu, Pressure pattern recognition in buildings using an unsupervised machine-learning algorithm, J. Wind Eng. Ind.
Aerod. 214 (2021), 104629.
[150] G. Hu, L. Liu, D. Tao, J. Song, K.T. Tse, K.C.S. Kwok, Deep learning-based investigation of wind pressures on tall building under interference effects, J. Wind
Eng. Ind. Aerod. 201 (2020), 104138.
[151] B. Kim, D.-E. Lee, K.R.S. Preethaa, G. Hu, Y. Natarajan, K.C.S. Kwok, Predicting wind flow around buildings using deep learning, J. Wind Eng. Ind. Aerod. 219
(2021), 104820.
[152] F.B. Chen, X.L. Wang, X. Li, Z.R. Shu, K. Zhou, Prediction of wind pressures on tall buildings using wavelet neural network, J. Build. Eng. 46 (2022), 103674.
[153] P. Lin, F. Ding, G. Hu, C. Li, Y. Xiao, K.T. Tse, K.C.S. Kwok, A. Kareem, Machine learning-enabled estimation of crosswind load effect on tall buildings, J. Wind
Eng. Ind. Aerod. 220 (2022), 104860.

24

You might also like