1 s2.0 S0038092X21004734 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Solar Energy 224 (2021) 516–530

Contents lists available at ScienceDirect

Solar Energy
journal homepage: www.elsevier.com/locate/solener

Early stage design for an institutional net zero energy archetype building.
Part 1: Methodology, form and sensitivity analysis
Samson Yip *, Andreas K. Athienitis, Bruno Lee
Building, Civil and Environmental Engineering, Concordia University, 1455 de Maisonneuve Boulevard West, Montreal, QC H3G 1M8, Canada

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

Keywords: This paper is the first of two that examine the influence of building form, including courtyards, on the energy
Net-zero energy building performance of medium-sized institutional solar net zero energy buildings incorporating building-integrated
Building form photovoltaic/thermal (BIPV/T) roof systems. The inclusion of BIPV/T at the early design stage along with
Building-integrated photovoltaic and thermal
passive design techniques extends to energy flexibility considerations with a smart grid. A two storey 3 000 m2
(BIPV/T)
Windows
medium-sized archetype building was considered. Through simulations, different building plan shapes and BIPV/
Sensitivity analysis T roof tilt angles supported by a range of related building enclosure variables such as insulation level and
Early stage design window characteristics were analyzed and compared. The design configurations used BIPV/T heat recovery for
winter ventilation air preheating in the Montreal region.
In this first paper, the methodology is presented. Then, using annual net energy use intensity as the target, a
global sensitivity analysis was performed to rank and quantify the most sensitive design variables at different
BIPV/T roof slopes. Out of the nine input variables analyzed, the most influential by rank were the orientation,
the window to wall ratio for the south and north façades, and the plan shape, for roof tilt angles of 35◦ , 40◦ , and
45◦ . For a roof tilt angle of 30◦ , the window to wall ratio of the south façade was the most influential, switching
rank with the orientation. Although not an input variable to the sensitivity analysis, this indicates as well the
influence of the BIPV/T roof tilt angle. Also, there was a significant interaction effect between the plan shape and
orientation, highlighting the utility of analyzing interaction effects at the early design stage and the importance
of these two factors in solar-enhanced buildings. At the other end of the spectrum, the window type, the wall and
roof insulation level, and the window to wall ratios for the east and west facades were the least significant – at
least approximately one order of magnitude less sensitive than the first ranked input.
In the second paper, the methodology is applied in a case study for early stage design in which relative
comparisons are made considering building form, net-zero energy, and extended to the grid considering load
matching and energy flexibility.

A common RET used in NZEBs is photovoltaic (PV) technology


(Athienitis and O’Brien, 2015). PV modules are often mounted on
1. Introduction building facades and roofs using their own racking structures to form so-
called building applied PV (BAPV). The PV can also be integrated as a
1.1. Solar net-zero energy buildings component of the building enclosure; creating building-integrated PV
(BIPV). In a BIPV system, the PV system is an integral part of the building
Renewable energy technologies (RETs) are being increasingly enclosure and follows the building form and contributes to the essential
adopted in the building sector as potential solutions to address the ur­ enclosure functions of support; access and circulation; finish; and con­
gent need for energy use reduction. They provide the energy-generating trol of rain, air, vapour, and heat (Straube and Burnett, 2005). An
component in net zero energy buildings (NZEBs). Although there are enhancement of BIPV is the building-integrated PV and thermal (BIPV/
different definitions and calculation methods (Marszal et al., 2011; T) system. This system increases the performance of a BIPV system by
Peterson et al., 2015), the strictest definition of an NZEB is a building using a fluid channel behind the PV modules for heat recovery (Athie­
that generates at least as much energy on-site from renewable energy nitis and O’Brien, 2015; Athienitis et al., 2015). The heat recovery
sources as it consumes in a year (Torcellini et al., 2006).

* Corresponding author.
E-mail addresses: samson.yip@mail.concordia.ca (S. Yip), andreask.athienitis@concordia.ca (A.K. Athienitis), bruno.lee@concordia.ca (B. Lee).

https://doi.org/10.1016/j.solener.2021.05.091
Received 19 October 2019; Received in revised form 14 March 2021; Accepted 25 May 2021
Available online 19 June 2021
0038-092X/© 2021 International Solar Energy Society. Published by Elsevier Ltd. All rights reserved.
S. Yip et al. Solar Energy 224 (2021) 516–530

Nomenclature WinID window identification number from LBNL Window


software
Symbols WWR window to wall ratio
ACH air changes per hour
CO Courtyard plan shape Greek symbols
COP coefficient of performance α wing rotation angle for L and U plan shapes [◦ ]
Edel energy delivered to the building site β tilt angle from the horizontal plane [◦ ]
Eexp on-site generated renewable energy exported from the κk collinearity index for kth term
building site πk sensitivity or contribution index for kth term
Elight electrical energy required for electric lighting Ψ azimuth angle, due South is 0; +ve clockwise [◦ ]
Eplug electrical energy required for all plug loads Acronyms
EPV electrical energy generated by the BIPV/T system exported ANOVA analysis of variance
from the building site BAPV building-applied photovoltaic
Esite net annual site energy BIPV building-integrated photovoltaic
Esource net annual source energy BIPV/T building-integrated photovoltaic and thermal
EUI energy use intensity DHW domestic hot water
HDD18 heating degree days below 18 ◦ C FAST Fourier amplitude sensitivity test
L L-shaped plan LBNL Lawrence Berkeley National Laboratory
PEF primary energy factor NECB National Energy Code of Canada for Buildings
Qcool required cooling energy NZE net-zero energy
Qheat required heating energy NZEB net-zero energy building
rdel source energy conversion factor for the delivered OFAT one factor at a time
electricity PV photovoltaic
rexp source energy conversion factor for the exported on-site RES renewable energy system
electricity generated from renewables RET renewable energy technology
RE rectangle plan shape SA sensitivity analysis
U U-shaped plan

assists in keeping PV panels cool, thus increasing their electricity pro­ optimization (Ourghi et al., 2007; Tuhus-Dubrow and Krarti, 2010). In
duction. The heat recovered can then be used in the building for low- this archetypal approach, simple displacement of vertices, or scaling of
grade processes such as ventilation air preheating (Athienitis et al., edges, wings, or aspect ratios (Tuhus-Dubrow and Krarti, 2010) creates
2011), space conditioning through a heat pump (Kamel and Fung, 2014) new forms. Further form variations have included wing rotations of the
and domestic hot water (DHW) heating (Kazanci et al., 2014). alphabet plan shapes in a parametric study of solar potential (Hachem
The high degree of functional integration in BIPV/T allows the po­ et al., 2011).
tential for envelope systems prefabrication and modularization, leading Different approaches concentrate on a singular base rectangle plan
to better quality and quality control, construction sequencing, materials shape or, by extrusion, the cuboid form and parameterize aspect ratios to
savings, and architectural integration. Due to this inextricable connec­ create variations. Hemsath and Alagheband Bandhosseini (2015) used
tion, a BIPV/T system’s electricity generating performance and the this method to optimize form based on energy consumption. Multiple
overall energy performance of the building will be influenced by the rectangle units were added and subtracted with different adjacency or
building form. overlapping conditions to create composite forms to optimize passive
Since building form is an important consideration in the overall design (Konis et al., 2016). A cuboid form’s vertices or surfaces were
design of a NZEB, it is essential to quantify its influence in attaining net parameterized and manipulated to optimize through genetic algorithm
zero energy (NZE) along with typical ‘semantic’ parameters such as the building form for minimum energy consumption (Yi and Malkawi,
thermal resistance, and thermal mass among others (Pessenlehner and 2009; Youssef et al., 2016). However, this strategy does not provide a
Mahdavi, 2003). This is particularly the case for medium-sized com­ large enough cross-section of possible building form.
mercial and institutional buildings – which account for the largest per­ Parametric scripting refers to the generation of building form
centage of floor space (33%) and energy use (31%) among all through computer scripting while maintaining dynamic links between
commercial and institutional building sizes in Canada (Natural Re­ parameters (Nembrini et al., 2014). This can be simple as in most of the
sources Canada, 2012) – where the largest surfaces for BIPV/T inte­ previous examples, but can become complex in investigations grouped
gration are roofs. around free-form buildings where the objective is form finding or form
exploration, like in Yi and Malkawi (2009). Key parameters can repre­
1.2. Building form diversity sent elements on various topological levels such as vertices, edges, and
faces. This is often paired with evolutionary algorithms in optimization
Martin and March (1972) proposed a set of formal building arche­ studies. Wang et al., (2006) generated polygonal forms through edge
types to simplify the analysis of urban building morphology and daylight length and angle or bearing representations to study lifecycle cost and
access: pavilions, slabs, terraces, terrace-courts, pavilion-courts, and life cycle environmental impact in a genetic algorithm optimization
courts. This work was seminal in introducing archetypes as a method of procedure. Other parameterizations include polygon triangulation and
building form analysis. Since then, they have been used for solar radi­ vertex manipulation (Kämpf and Robinson, 2010) and surface repre­
ation analyses for shadow density, daylight distribution, sky view, and sentation by Fourier series, Taylor series, snake forms, and squared
maximizing incident solar radiation (Kämpf et al., 2010; Ratti et al., forms to optimize or minimize solar irradiation on façade and roof using
2003). Others have used alphabet shapes (like L, and T, among others), an evolutionary algorithm (Caruso and Kämpf, 2015; Kämpf and Rob­
cross, rectangle, and trapezoid plan shapes as the basis for creating a inson, 2010). Non-uniform rational B-splines (NURBS) and control
variety of building forms to study energy consumption through curves were used to generate the free form of a community center

517
S. Yip et al. Solar Energy 224 (2021) 516–530

building in a genetic algorithm optimization for solar radiation, shape determine how much of the total variances of the output have been
coefficient, and space efficiency (Zhang et al., 2016). accounted for in the analysis.
Shape grammars can be considered separately as generative design In global SA methods, the input parameters are all varied together
systems. They use a set of conditional, ‘if-then,’ rules to create building over the entire range of the input space. The advantages of global SA
forms that follow coded compositional principles. They are difficult to methods include the ability to capture the total variations in the output
create due to their complexity and the results are idiosyncratic – sty­ and the effects of input parameter interactions. However, this comes at
listic, like Frank Lloyd Wright’s prairie houses (Granadeiro et al., 2013). the cost of greatly increased computational time. The most common
The previous works mentioned don’t fall neatly into the same cate­ global methods used in building energy analysis are based on regression,
gories across: approach to and type of building form; building perfor­ screening, and variance. The regression models like standardized
mance objective and methodology; building function and size; and regression coefficients are easy to implement and understand, and fast to
inclusion of BIPV. Most previous work concentrated on single-family compute, but are unsuitable for non-monotonic models, and depending
dwellings with a smattering of high-rise commercial (Ourghi et al., on the method, non-linear models (Tian, 2013).
2007; Youssef et al., 2016), and medium-sized institutional/office The Morris method is a screening method. It is a qualitative measure
(Konis et al., 2016; Zhang et al., 2016). Most analyzed daylighting and/ for ranking important input parameters from a large number of input
or energy consumption, but the single-family housing results cannot be variables without reducing the output variance. It requires relatively
extrapolated to larger institutional buildings due to differences in scale, few simulations per input parameter compared to other SA types; it can
occupant density, occupancy schedules, and internal loads, among handle non-linear models and correlated inputs; it uses discrete values of
others, which will affect building physics differently. BIPV was consid­ the input parameters instead of probability distributions; it traverses the
ered in the case of small scale domestic architecture (Hachem et al., entire input space; its output sensitivity indices are easy to interpret and
2011; Kämpf and Robinson, 2010) and high rise commercial (Youssef can show non-linear effects and input correlations (Heiselberg et al.,
et al., 2016); neither are applicable to the archetypes of medium-sized 2009; Nembrini et al., 2014; Tian, 2013).
institutional/commercial buildings due to different sizes and roof The following papers used the Morris method. Nembrini et al.,
placement of BIPV. Finally, only a couple studies featured the courtyard (2014) used an integrated dynamic model (Negendahl, 2015) to study
shape (Konis et al., 2016; Ratti et al., 2003) in spite of its advantages the influence of window to wall ratios (WWRs) and buffer spaces on
such as daylight distribution and microclimate (Ratti et al., 2003), of comfort frequency of occupants and energy demand for a high-rise
benefit even in cold climates (Martin and March, 1972). However, those apartment building. However, they did not change the building form.
courtyard studies only focused on passive design. Semantic variables Hemsath and Alagheband Bandhosseini (2015) analyzed a single zone
like U-values, window properties, etc. were studied in parallel with residential building using rectangular plan aspect ratio, stacking height,
building form for all of the studies cited, but while the optimization orientation, WWR and semantic variables such as window U-value and
studies’ objectives were form finding or form optimization, the para­ wall insulation R-value for different cities to target pre-design heating
metric studies looked for broader influential form relationships as design and cooling energy performance, but did not have a separate WWR for
guidance. each façade orientation, nor did they study different plan shapes. Hei­
selberg et al., (2009) investigated 21 input parameters’ influence on
1.3. Sensitivity analysis energy use reduction for an office building, however the building form
was not varied. (Garcia Sanchez et al., 2014) investigated 24 design
The choices of, and constraints to, building form, materials, all the variables in an apartment building including the length, width, height,
way to siting and orientation may be the result of factors outside of the and separate glazing ratios of the facades for influence on yearly heating
realm of net energy design. Parametric analyses are typically used loads, among others, but did not study other building forms.
(Hachem et al., 2011) to assess the impacts of changing input design The Analysis of Variance (ANOVA) method decomposes the output
parameter values one at a time. The limitations of such a process become variance to account for the contribution of each input factor. It considers
evident once the number of input design variables becomes large. To main effects, representing the impact of the uncertainty of the input
assist in such design support, sensitivity analysis (SA) can be used. SA factors on the output variance; second order effects, representing the
relates the variations or uncertainty in a model output due to the vari­ interaction effects of two input factors on the output variance; and
ations or uncertainty in the inputs. Benefits for building design include higher order effects, representing the interaction of higher order
model simplification from screening analysis, quantification of sensi­ numbers of input factors on the output variance. It is considered a
tivity, and identification of input variable interactions (Hopfe and model-free approach because it can account for complex nonlinear, non-
Hensen, 2011; Tian, 2013). By using SA as part of an integrated design additive, and non-monotonic models. Mechri et al. (2010) used ANOVA
process, identifying such variables early on gives the design team the on an intermediate floor of a rectangular-shaped plan office building and
greatest chance to respond to such information and improve the design. found that the envelope transparent surface ratio and the compactness
Research has shown that design changes made early in the design pro­ ratio (surface area to volume) were the most influential parameters on
cess have less adverse time or budget consequences than in later design energy performance. The windows were distributed uniformly across all
stages (Paulson, 1976). facades but no other building plan shapes were studied.
Different SA types are used based on the nature and quantity of the Lam et al. (2016) used ANOVA to study the influence of ten design
input parameters, the model, the output required, and the analysis ob­ parameters for a BIPV curtain wall office perimeter zone on annual
jectives. The different types used in building energy analysis have been energy consumption and annual energy balance – the ratio of annual PV
reviewed by Tian (2013). SA can be grouped into two approaches: local electricity generation to annual energy consumption. Their findings
and global. Local methods – also called differential sensitivity analysis – showed that WWR, and U-glazing were sensitive for both energy con­
are typically focused on the effects of a single parameter on the model sumption and energy balance, while the infiltration rate was sensitive
output variation. The analyses often use a one-factor-at-a-time (OFAT) for energy consumption and PV efficiency for energy balance.
method in which the impact of the range of values of a single input Chen et al. (2019a, 2015) used a single zone apartment unit with
parameter is evaluated on the output, with all other parameters held limited natural ventilation and daylight on an intermediate floor of a
constant. The local approach can be useful to explore a restrained, local, 30–40 storey apartment tower to represent a worst-case design scenario.
input space around a reference case. It is very easy to implement, has BIPV was integrated on the one façade on both opaque and fenestration
low computational cost, and the results are easy to interpret. However, areas to create a nearly-zero energy building model for a combined SA
the limitations of the local approach include the inability to identify and optimization process. The SA used the Morris and the variance-
interactions or correlations between parameters and the inability to based Fourier Amplitude Sensitivity Test (FAST) method on 11 input

518
S. Yip et al. Solar Energy 224 (2021) 516–530

variables for the target of net energy demand. For the FAST total order temperature, the temperature of the outdoor air, and the set point
index and the Morris method with sufficient sampling, building temperature. This is a highly efficient heat recovery system because
airtightness was the most influential input variable, followed by window there are no heat exchangers involved. This system was used in the
to floor area ratio, visible transmittance, and light to solar gain (LSG) Varennes Library (Dermardiros et al., 2019) and the John Molson School
ratio. Building orientation was ranked sixth. of Business at Concordia University (Athienitis et al., 2011).
Li et al. (2018) also used an integrated SA and optimization approach The BIPV/T outlet air can be used as input to an air source heat pump
to reduce total annual energy consumption and winter thermal in heating mode. The efficiency of a heat pump decreases with
discomfort for subtropical buildings with only cooling systems. Their decreasing outdoor temperature. At really low temperatures, this effi­
case study building had a BIPV-clad roof and a mix of commercial and ciency is further decreased by defrost cycles triggered by freezing
residential uses. Of the 29 input factors, 14 were found to be sensitive evaporator coils from condensation. When a BIPV/T system is gener­
across regression, Morris, and FAST SA methods; among them were the ating electricity, its warm outlet air can be fed over the evaporator coils
skylight to roof ratio, WWR, and building orientation. Six, including of a heat pump to increase the source air temperature and hence the
orientation and WWR, were then chosen for optimization. The building efficiency of the heat pump. Kamel and Fung (2014) performed a
shape was fixed and there was only one WWR for all facades. simulation study for a roof-integrated BIPV/T system coupled with an
Chen et al. (2019b) incorporated SA as part of an optimization air source heat pump. They found that the combination could lead to
approach for a 40 storey high-rise office building with BIPV facades. electricity cost savings and greenhouse gas emissions reduction. The
Using the Morris and variance-based FAST methods with net energy BIPV/T outlet air has also been transferred through an air-to-water heat
demand as the target, they found the WWR, visible transmittance (VT), exchanger to a water-source heat pump coupled with a storage tank for
LSG, and window U-value to be the most significant. Separate SAs with space heating in an experimental and simulation study (Nejma et al.,
square, rectangle, H-, U-, T-, and L-shaped plans did not show any dif­ 2013) and through an air-to-water heat exchanger directly to a domestic
ference in the ranking of the significant inputs. However, the optimi­ hot water tank (Chen et al., 2010). Chen et al. (2010) also thermally
zation showed that the different plan shapes significantly influenced the connected the BIPV/T system to a ventilated concrete slab for space
energy performance. The optimal square shape design had the least net heating. Another use of the heat is through an air-to-water heat pump for
energy demand reduction at 36.83%, and the optimal H-shape design DHW and space heating (Delisle and Kummert, 2016).
had the most at 48.77%, as compared to their respective reference cases. There are few studies that relate building form diversity to BIPV/T
Using the shape coefficient (SC), the ratio of envelope area to condi­ implementation. Previous studies including passive and active energy
tioned volume, they concluded that increasing SC increased both HVAC effects, and BIPV/T energy generation have focused on small-scale res­
demand and PV generation. The same WWR variable was used for all the idential roof applications (Delisle and Kummert, 2016), office or insti­
facades, the roof area was not a significant element in their study, and tutional high-rise BIPV/T façade applications (Athienitis et al., 2011,
plan shape was not a factor in the SA. 2018). In the case of small-scale (Buonomano et al., 2016) and medium-
The previous SA works have included common early design stage sized (Dermardiros et al., 2019) institutional BIPV/T roof applications,
factors such as component U-values and window properties. Some have net energy use was assessed without building form analysis or variation.
identified non-linear and non-monotonic relationships between the
factors and target such as WWR, orientation, and thermal insulation 1.5. Research needs
(Chen et al., 2019b; Garcia Sanchez et al., 2014), making variance-based
SA approaches suitable for such models. However, only one defined As performance expectations increase for high performance build­
separate WWR factors for each façade orientation to allow WWR ings such as NZEBs, so does the need for building systems integration
behaviour to be analyzed by façade (Garcia Sanchez et al., 2014), but and the need for research at the intersection of the disciplines that
did not include renewables. Only one SA study included a medium-sized collaborate in building design. The previous sections have shown there
institutional building with roof-integrated BIPV system, but did not are few studies for medium-sized NZEBs – the medium-sized institu­
investigate other building plan shapes (Li et al., 2018). Other SA papers tional building type being a significant component of Canadian city­
that did have building form variation addressed it as a series of aspect scapes – with BIPV or BIPV/T roofs and none of those have examined
ratios or proportions of a reference form: directly, in plan and section building form variations. In general, the studies that included building
through aspect ratios of a single-family dwelling (Hemsath and Ala­ form as a variable tended to evaluate either passive or active design – but
gheband Bandhosseini, 2015); or indirectly, abstracted as envelope area seldom both like in a NZEB; those that did used high-rise buildings with
to volume ratios of an intermediate floor plate of an office building façade-integrated PV. And the forms tended not to include the courtyard
(Mechri et al., 2010). This does not provide enough of a variety of type – an important archetype in the urban fabric. Finally, of those few
building form to generalize results. Finally, only one SA study featured a papers which analyzed the influence of building form on net energy
variety of plan shapes, but the building type was a high-rise with façade- consumption, none have quantified the contribution of building form as
integrated BIPV – which poses different geometry design challenges than a factor on net energy consumption variance in a sensitivity analysis.
medium-sized roof-integrated BIPV – and it was found that plan shape In light of this, the objective of this paper is the integration of these
did not influence the ranking of the significant inputs (Chen et al., aspects: to quantify the influence of building form, including courtyards,
2019b). However, plan shape was not included as a factor in the SA, passive and active design performance, and BIPV/T integration in
therefore its significance compared to the other factors such as WWR is medium-sized institutional NZEBs through sensitivity analysis.
not known. First, the approach of archetypical plans – including the courtyard –
was extended to and developed as medium-sized institutional NZEBs
1.4. BIPV/T and integration of systems with building design with a BIPV/T roof system to represent a range of possible building
forms. These were parameterized in an automated workflow. The ar­
There are different ways to utilize the heat recovery from a BIPV/T chetypes were then evaluated in a global sensitivity analysis where plan
air-based system in a cold climate. The simplest is for ventilation air shape was included as a factor – among other typical early design stage
preheating. Outdoor air is always required for the operation of a variables – to quantify its influence on NZE.
building. During winter this increases the heating load on the ventilation All told, the interactions between building plan shape, BIPV/T roof
system when heating outdoor air to required supply temperature. Pre­ area and tilt angle, orientation, window to wall ratios, window prop­
heating some or all of this air through the BIPVT/T system can reduce erties, and thermal insulation and their effects on the combined passive
this load considerably. The BIPV/T outlet air supply quantity can be performance, PV electricity generation, and heat recovery were evalu­
controlled proportionally based on the difference between its ated through an ANOVA global sensitivity analysis targeting annual net

519
S. Yip et al. Solar Energy 224 (2021) 516–530

energy use intensity (EUI). The authors are not aware of a parametric storeys in order to include interactions between vertically adjacent
study or sensitivity analysis in the literature that has addressed this set of spaces. Each floor was divided into four wings of 375 m2 net floor area
design variables and objectives together. each in an open-plan type of space. The plan shape of each of the wings
A cold climate with significant cooling loads was used for the sim­ may vary but the total floor area was the same to permit comparison
ulations. The heat recovered from the BIPV/T system was used for across different plan shapes. The two storeys were stacked; the option to
winter ventilation air preheating. The archetypes created were partly have different floor areas for each of the storeys was not studied, nor was
inspired by the Varennes Library, on Montreal’s South shore in Var­ the possibility of non-vertical facades. The height of each storey from the
ennes, Quebec, Canada, which is Canada’s first net zero energy insti­ finish floor to the underside of the deck was 3.5 m. This did not include
tutional building. It is a BIPV and BIPV/T roof-clad building of 2 storeys the ceiling space above the 3.5 m height on the upper storey which was
and 2 100 m2. The first year operational site EUI was 70 kWh/m2a; an additional triangular volume that was created by the pitched roof.
including renewables, the net energy use intensity was 14.5 kWh/m2a. The building design included 200 mm of concrete in the envelope wall
Alternatively, applying the relevant primary energy factor for source composition as well as exposed on the floor serving as thermal mass.
energy, the building reaches net zero energy (Dermardiros et al., 2019). The depth of a wing was the dimension measured perpendicular to
A detailed methodology describing the archetypes and modeling the longest façade. It was determined based on the goal of enhanced
assumptions is contained in the next chapter followed by the sensitivity daylighting permitted by the open-plan space type. Previous studies
analysis findings and conclusion. have determined that there is a limiting depth for useful daylighting
In the second paper, the methodology is applied to an early stage (Guglielmetti et al., 2010; Yip et al., 2015). For this paper, the depth was
design case study for annual net-zero energy, supported by the results of set at 15 m.
the sensitivity analysis. Early design stage input variables in design The plan shapes, shown in Fig. 1, were inspired in part by those
configurations are compared on a relative basis and the configurations derived in earlier work by Martin and March (1972) who decomposed
are then analyzed in the context of the grid: the impacts on load urban built form into a series of reference shapes. These plan shapes
matching and energy flexibility. represent individually, or built up in combination, common plan shapes
found in urban settings. The rectangle plan shape, RE, was the reference,
2. Methodology and had its long axis oriented east–west. The L- and U-shapes were
further defined by a plan rotation angle from 15◦ to 75◦ towards the
2.1. Building design components and assumptions north in a way that maintained solar exposure while minimizing self-
shading for the inclined BIPV/T roof. The courtyard, CO, shape was
Based on a national survey of commercial and institutional buildings the only plan shape with any significant self-shading. It offered a central
in Canada, the medium-sized building (930–4 645 m2) is the category courtyard that was semi-enclosed, being open to the sky. The courtyard
with the second largest number of buildings after the very small building may be further protected with a roof covering; however transforming
(465 m2 or less) and with the largest percentage of floor space (33%) and this courtyard into a semi-conditioned or interior space would have
energy use (31%) among the different building sizes in the country increased the total floor area of the building making it unequal in floor
(Natural Resources Canada, 2012). Buildings in this size category are area to the other plan shapes and turning it into a cubic building. The
generally low-rise, of a few storeys tall at most, making the roof the L15 shape, in Fig. 2, has the closest resemblance to the plan shape of the
largest enclosure surface on which to integrate a BIPV/T system. Varennes Library.
Therefore, this study used the equator facing roof as the BIPV/T surface The roof was pitched symmetrically about the longitudinal axis of
to maximize electricity and heat production so as to reach or exceed the each wing. In order to maximize solar radiation capture, the BIPV/T
net-zero energy target. system was integrated on all roof surfaces between azimuth angles of
− 90◦ and 90◦ . Where the longitudinal axis was oriented north–south,
2.1.1. Climate both the east- and west-facing roof surfaces were clad with a BIPV/T
The climate of Montreal, Quebec, Canada, Table 1, was used for the system. This was the case for the courtyard shape, CO, allowing for two
simulations. The minimum building enclosure thermal requirements more PV roof surfaces than for the other shapes. The gables, which are
were based on the National Energy Code of Canada for Buildings (NECB) the triangular surfaces between the two intersecting roof pitches, were
(Canadian Commission on Building and Fire Codes and Construction, vertical and not used for BIPV/T.
2015). For the medium-sized building with roof-integrated BIPV/T typol­
ogy, the building depth limitation for daylighting is interconnected with
2.1.2. Functional and technical program the BIPV/T roof geometry. A larger building depth will lead to a taller
This paper focused on open-plan spaces in commercial/institutional roof, if respecting the same BIPV/T tilt angle, as shown in Fig. 3, which
buildings such as for office, or informal assembly uses such as libraries. creates large interior volumes that may not necessarily be useful. This
Open-plan spaces permit daylight to penetrate deeper into spaces to save triangular volume can be partially or completely closed to form
on electric lighting costs and enhances natural ventilation to save on mezzanine or attic spaces. However, this would add additional floor
cooling costs. space, exceeding the 3000 m2 of net floor area that is a constraint for this
paper. Additionally, this extra net floor area is not without compromise:
2.1.3. Building archetype the walls would be inclined inwards with no fenestration on the south
The two components of building form studied were plan shape, and side. It would require detailed architectural design to fulfill a specific
BIPV/T roof type and tilt angle. A medium-sized building with a total function unencumbered by the mezzanine or attic area restrictions. And,
useable, net floor area of 3000 m2 was used. The building had two it adds to the material quantities and costs of construction. This can to be
studied in the future as an optimization problem. For example, in the
Varennes Library, the BIPV/T roof pitch is 37◦ , a compromise between
Table 1
solar radiation capture, limiting roof height and volume, and reducing
Selected climate characteristics for Montreal, Quebec, Canada (ASHRAE, 2016;
snow accumulation in winter.
Government of Canada, 2019). HDD18 are heating degree days below 18 ◦ C.
Triangular roof surfaces are inevitable when intersecting different
City Lat/Long Zone ASHRAE Annual sunshine geometrical configurations and carry inherent inefficiencies when using
HDD18 hours
different roof shapes as the surface for BIPV and BIPV/T design appli­
Montreal, 45.5◦ N, 6 4603 2051 cations (O’Brien et al., 2015). This inefficiency was acknowledged and
QC 73.58◦ W
accounted for in the building model. An example of the net BIPV/T roof

520
S. Yip et al. Solar Energy 224 (2021) 516–530

Fig. 1. Plan shape families; grey shading indicates BIPV/T roof surfaces; L- and U-shapes show wing rotation angle (α). Linear measurement in millimeters.

total floor area, the CO shape has more BIPV/T surface area but at a
reduced 85% roof surface area utilisation efficiency.

2.1.4. Renewable energy technologies: BIPV/T


The BIPV/T system was an integral part of the south- and near south-
facing roof of the building. A nominal PV electrical conversion efficiency
of 17% was used across all simulations. For the heat recovery in the
BIPV/T system, this paper assumed an open-loop air-based system. A
schematic is shown in Fig. 3. While this system may be less efficient than
a hydronic system, an air-based system is easier to install in modular
construction, requires little maintenance, tolerates sub-zero tempera­
tures, and has less risk of causing building damage in the case of a

Fig. 2. Legend; plan shape number in parentheses. Grey shading indicates


BIPV/T roof surfaces.

surface area compared to gross roof area available is shown in Fig. 4. For
example, at a 35◦ roof tilt angle, the RE and CO shapes each have an
overall gross roof surface area of 1831 m2. For the RE shape, half (915.5
m2) of this is sun-facing and available to receive BIPV/T. Since the
rectangular roof would be 100% efficient for standard rectangular BIPV/
T deployment, the net BIPV/T surface area would be the same
(neglecting PV framing and other such elements). The CO shape has Fig. 4. 3D view of roof volume showing BIPV/T surface in blue and residual
1373 m2 gross roof surface area available to receive BIPV/T, but the net roof area in light grey. (For interpretation of the references to colour in this
BIPV/T surface area is 1167 m2. Thus while all shapes have the same figure legend, the reader is referred to the web version of this article.)

Fig. 3. Building section showing roof tilt angle (β) and height (in millimeters) and detail of BIPV/T system.

521
S. Yip et al. Solar Energy 224 (2021) 516–530

component or system failure as compared to a liquid system. The height Table 3


of the air channel in the BIPV/T system was fixed at 25 mm. While this Input variables: window properties. WinID is the window ID; the U-value and the
parameter may be varied, it was assumed that varying the height had the SHGC are the overall values for the window; SHGC is the solar heat gain coef­
same, if not identical, effect across all building configurations. A single ficient; Tsol is the solar transmittance; Tvis-daylight is the visible light
inlet system was used; multiple inlet systems show promising perfor­ transmittance.
mance improvements (Rounis et al., 2016; Yang and Athienitis, 2015), Parameter Window 1 Window 2 Window 3
but are beyond the scope of this paper. Description Double-glazed, Double-glazed, Double-glazed,
Due to the large sample size of design configurations studied, details air-filled, low-e argon-filled, low-e krypton-filled, low-e
of the PV/T system such as the exact specifications of the PV modules, WinID 3200 3205 3210
U-value (W/ 1.69 1.40 0.97
the array configuration and other balance of system components may
m2K)
vary significantly and were not examined in this paper. These simpli­ SHGC 0.66 0.70 0.53
fying assumptions are still useful at the early design stage for making Tsol 0.562 0.555 0.445
relative comparisons. But it is acknowledged that as a building design Tvis-daylight 0.80 0.74 0.70
moves from early stage design into design development, there may be
details of the PV system that emerge that would benefit some design
options and penalize others. Table 4
Non-varying parameters.
2.1.5. Related geometrical and other parameters Parameter Value(s)
Window to wall ratio (WWR): A separate WWR was calculated for
Floor plate (m2)/storeys 1 500/2
each building face orientation. The WWR has an influence on
Total floor area (m2) 3 000
daylighting effectiveness, and solar thermal gains. Concrete
Window types: The windows were composed of three components, Density (kg/m3) 2 240
the glazing, the framing, and the shading system. The glazing was Exposed on floor, thickness (m) 0.2
selected from the LBNL Window v7.4.6.0 database (LBNL, 2016). LBNL Within exterior wall assembly, thickness (m) 0.2
BIPV/T channel height (m) 0.025
Window is a software program that calculates total window thermal PV nominal electrical conversion efficiency 0.17
indices such as: U-value; solar heat gain coefficient (SHGC), the fraction Heating COP 2.5
of the incident radiation transmitted and absorbed and then radiated Cooling COP 3.5
inwards; solar transmittance (Tsol), the fraction of the incident radiation Heating Setpoint/Setback (◦ C) 21/18
Cooling Setpoint (◦ C) 24
transmitted from the entire solar spectrum; and visible transmittance
Schedule (occupancy) 07:00–20:00
(Tvis-daylight), the fraction of the incident radiation transmitted from Occupant density (m2/person) 10
the visible spectrum (380–700 nm). Users can create windows from any Occupant gains (W/m2) 7.5
combination of glazing layers, gas layers, frames, spacers, and dividers All humidity gains (kg/h*m2) 0.016
using their own components or chosen from an included database of Plug loads (W/m2) 9
Electric lighting (W/m2) 7
commercial components. As described in Section 2.1.1, glazing units Infiltration (ACH @ 50 Pa) 1*
that satisfied the minimum NECB 2015 requirements were used in this Ventilation (L/s*person) 5.6
paper. The framing accounted for 15% of the window (2.27 W/m2 K)
*The infiltration at natural pressure was estimated as 0.05 ACH using the K-P
and there was an automatically controlled interior solar shading system
model (Sherman, 1987).
on all windows except those on the north façade.
Insulation: The minimum levels of wall insulation and roof insulation
(Klein et al., 2017). It is a widely used platform in building energy
were selected based on NECB 2015 requirements. The maximum insu­
performance simulation research and has been the subject of numerous
lation values were determined based on a case study of a Passivhaus
validation studies (Crawley et al., 2008). Its structure is modular,
building for the Montreal climate.
allowing users and developers alike to custom tailor or restrict their
Details of input variables and non-varying parameters are shown in
efforts to specific components of an energy system without needing to
Tables 2–4.
engage with all the other components of the TRNSYS software. The
components are referred to in TRNSYS as ‘types.’ Each type simulates a
2.2. Simulation model and assumptions component or sub-component of a building or energy system and re­
quires specific inputs to produce outputs. There can be numerous in­
2.2.1. TRNSYS model stances of each type in a model. The manner in which the units are
The building performance simulation program used was TRNSYS 18 connected via their inputs and outputs allows a high degree of functional
flexibility and customization of building design configurations. The
Table 2
configuration of all these types and their connections are contained in
Input variables. the TRNSYS .dck file.
The principal types used in this project are the TYPE 56 which rep­
Input variable Range of values
resents a multi-zone building and TYPE 568 which represents the
Plan shape RE, CO, L15, L30, L45, L60, L75, U15, U30, U45, building-integrated PV and thermal system. The TYPE 56 manages the
U60, U75
3d geometry and windows to describe a building, and has settings for
BIPV/T tilt angle (◦ ) 30–45; 4 levels
Azimuth (◦ ) − 45 to 45; 7 levels building environmental controls, occupancy schedules, and for calcu­
Roof insulation (m2K/W) 5.7–12.4; 5 levels lating ideal heating and cooling loads. This information is contained in
Wall insulation (m2K/W) 3.9–10.5; 5 levels the TRNSYS .b18 file which is referenced in the .dck file. Each unique
Window See Table 3 building combination of plan shape and roof tilt angle required a
WWR; south, east, west, 0.1–0.7; 7 levels each orientation
north
separate TYPE 56 instance. Detailed models of components such as
Dependent variables Range of values mechanical systems are contained in other types that connect to TYPE
56.
Total roof surface area (m2) 827–1591
Total BIPV/T surface area 827–1233 An on/off control signal contained in TYPE 56′ s built-in ‘Daylight
(m2) Control Type Manager’ was used to turn the electric lighting on or off

522
S. Yip et al. Solar Energy 224 (2021) 516–530

based on two defined horizontal illuminance setpoints of 300 lx and 500 was devised (Fig. 5). An initial set of four building energy models, one
lx. An on/off shade control signal contained in TYPE 56′ s ‘Window Type each for the CO, L, RE, and U plan shape families, represented by four
Manager’ was used to open or close the interior shade based on total sets of TRNSYS .dck (building TYPEs and connections) and .b18
solar radiation setpoints of 120 W/m2 and 140 W/m2 impinging on the (building description file including 3d geometry) input files, was
associated window. Both controls use a hysteresis loop to turn the signal parameterized by replacing input values with variable expressions based
on or off based on the setpoints at each timestep. on wing rotation, α, (Fig. 1), roof tilt angle, β, (Fig. 3), and the input
In the present paper, the BIPV/T surface was the roof. In TRNSYS the variables listed in Table 2. These modified .dck and .b18 files served as
roof insulation properties were input to both the TYPE 56 and the TYPE base template files.
568 units. Then the TYPE 568 for the BIPV/T was connected to the TYPE The number of simulations required was input to the Sobol’ sampling
56 to obtain the inside surface temperature from the building zone and routine in modeFRONTIER (ESTECO SpA, 2017) which generated the
to transmit to the TYPE 56 the back face temperature of the BIPV/T air combination of values for each of the input parameters to create a
channel. Note that for well-insulated roofs, there will be little energy unique building design configuration to simulate. modeFRONTIER then
exchange between the BIPV/T air channel and the building zone below passed these values to a Python (Python Software Foundation, 2019)
it. script which automatically generated a new building model including
Each two-storey building wing, totalling 750 m2 of floor area, was the required building roof pitch, wing rotation (where applicable), and
clad with at least one BIPV/T roof system. For the RE, L, and U shapes, WWR from the base template .dck and .b18 files and launched the
each BIPV/T system was 25 m wide and the heat recovered served only TRNSYS simulations. The output from each TRNSYS simulation was
the wing below it. For the courtyard building, the three outward facing collected in a database within modeFRONTIER. This procedure repeated
BIPV/T roof systems were also 25 m wide and served their respective until all the unique building configurations were simulated. Once all
wings while the three smaller 10 m wide BIPV/T roof systems facing the simulations were complete, results were sorted, visualized, and a
inner courtyard served the north wing. The outlet air speed of each sensitivity analysis was performed.
BIPV/T system was set to a constant 0.67 m/s, the lowest speed possible
that would maximize the heat recovery and supply the entire ventilation 2.2.3. Performance metric: Net zero energy
need of the wing it was meant to serve when required. This was based on Since this investigation was focused on the early stages of design
the channel width of 25 m and height of 25 mm, for a cross sectional when the major design decisions for a building are yet to be taken, some
flow area of 0.625 m2 per wing, and a maximum required ventilation simplifying assumptions were made. Specific mechanical systems were
rate per wing of 0.42 m3/s (or 5.6 L/s per person). The net channel width not studied and no energy losses due to system inefficiencies or in­
for the courtyard north wing was 30 m (three 10 m widths), making the teractions were calculated. Instead, an ideal heating and cooling system
volumetric flow rate slightly higher than required. was used to determine the heating and cooling demands of the building.
The BIPV/T outlet air contribution to the supply air was controlled To approximate a total consumption, a separate constant heating and
based on the difference between its temperature, the temperature of the cooling coefficient of performance (COP) of 2.5 and 3.5, respectively,
ambient air, and the set point temperature. TYPE 56 calculates any was assumed based on heat pump technology.
remaining energy required to bring the outdoor air to room conditions. The energy balance formulation used for the building designs was the
For increased PV module cooling and thus PV generation, a higher air annual import/export balance which is the balance between delivered
flow speed can be used in the BIPV/T air channel. For increased heat (demand) and exported (supply) energy (Sartori et al., 2012). The en­
recovery, a lower air flow speed with a taller air channel height can be ergy balance was calculated using net-zero site energy (Torcellini et al.,
used. Different BIPV/T roof shapes will also change this dynamic, as 2006). Since this was a single building with BIPV/T, the site boundary
would a variable air flow speed in the air channel. The balance between was the building footprint for the net-zero energy balance calculations.
PV cooling, heat recovery, and ventilation requirements can be fine- The building was all-electric, therefore only one energy carrier was
tuned after the early design stage or once specific equipment is considered for delivered energy.
selected to match specific project needs. The net annual site energy, Esite, using hourly timesteps, is given by
Further modeling detail was omitted to focus attention on using the Eq. (1). This is normalized by building size to obtain the net EUI by
model for comparing the relative performance of the different building dividing Eq. (1) by the building’s total conditioned floor area. Esite ≤
geometries at the early design stage. A more detailed analysis of the 0 indicates the building is a NZEB.
relationships between the PV electrical conversion efficiency, heat re­ ( )
Qheat Qcool
covery and air flowrate, ventilation, and BIPV/T roof geometry is a topic Esite = Edel − Eexp = + + Elight + Eplug − (EPV ) (1)
COPh COPc
for future investigation.
where Edel is the energy delivered to the building site;
2.2.2. Simulation workflow
To permit the automation of the simulations, the following workflow

Fig. 5. Flowchart of workflow; * the input variables are listed in Table 2.

523
S. Yip et al. Solar Energy 224 (2021) 516–530

Eexp is the on-site generated renewable energy exported from the f(x) with Gaussian-type response, the responses
building site;
yi = f (xi ) + ∊i , i = 1, ⋯, n, (4)
Qheat represents the heating energy required, including for ventila­
tion, and to compensate for infiltration losses; where n is the number of samples, and ∊i represent independent random
Qcool represents the cooling energy required, including for ventila­ errors. Estimation of f can be done with a smoothing spline, or mini­
tion, and to compensate for infiltration gains, and internal gains from mizer fλ , that satisfies the following penalized least squares minimiza­
occupants, lights, and equipment; tion function:
COPh and COPc are the heating and cooling COPs, respectively;
Elight is the electrical energy required for electric lighting; 1∑ n
min (fi − f (xi ))2 + λJ(f ), (5)
Eplug is the electrical energy required for all plug loads; and n i=1
EPV is the electrical energy generated by the BIPV/T system exported
from the building site. where λ is the Lagrange multiplier, or smoothing parameter, and J(f) is
the roughness penalty which serves to prevent overfitting f.
Note that in winter, the BIPV/T thermal energy recovered is used to The SS-ANOVA function decomposition can be incorporated into Eq.
supply the heating energy for preheating the ventilation air, as described (5) through the use of reproducing kernel Hilbert spaces (Gu, 2013).
in Section 2.2.1, thus reducing the need for delivered energy to the When evaluating the model fit at the sampling points xi , from Eq. (4), the
building. following is obtained as expressed using column vectors of length n:
While site energy has been used in this paper for simplicity in the all-
y = f 0 + f 1 + ⋯ + f p + e, (6)
electric building, a source energy definition is more generally applicable
and can allow comparisons of the energy performance of buildings using where f 0 is a constant and p is the number of tensor sum terms deter­
different energy carriers or across different jurisdictions. In that case, the mined by the number and type (main or interaction effect) of effect
energy delivered and exported would be weighted through the concept terms in the model. Further simplification by removing the constant
of source energy conversion factors (or primary energy factors) to ac­ term through projection (Gu, 2013) gives:
count for the energy consumed in the extraction, processing and trans­
port of primary fuels; losses in the generation process; and losses in y* = f * + e* , (7)
transmission and distribution to the building site (Peterson et al., 2015;
Voss and Musall, 2011). Furthermore, using the United States Depart­ where
ment of Energy (US DOE) formulation (Peterson et al., 2015), the PV p

exported electricity can be credited for displacing electricity required f* = f *k . (8)
from the grid, thus its source energy conversion factor would be equal to k=1

that of the delivered electricity. Therefore, using the US DOE source The relative contribution to the output variance of each term in the
energy definition for an all-electric building, Eq. (1) can be modified to model, the sensitivity or contribution index, πk , is
obtain the net annual source energy, Esource, as shown in Eq. (2).
(f *k , f * )
(
Qheat Qcool
) πk = , (9)
Esource = Edel rdel − Eexp rexp = + + Elight + Eplug rdel − (EPV )rexp ‖f * ‖2
COPh COPc
(2) where (u,v) is the scalar product of vectors, and ‖v‖ is the vector norm.
This shows the advantage of ANOVA: it provides a straightforward
where rdel and rexp are the source energy conversion factors for the interpretation of the significance of each term on the output variance.
delivered electricity and exported on-site electricity generated from A measure of the correlation of input variables can be calculated
renewables, respectively, and are equal. Under these conditions, Eq. (2) through a collinearity index, κk . Define a p × p matrix C such that
can be simplified using Eq. (1) to:
(f *i , f *j )
Esource = Esite rdel = Esite rexp Cij = , (10)
‖f *i ‖∙‖f *j ‖
2.2.4. Sensitivity analysis and sampling method
then
Since the purpose of this analysis was to identify at the early design
√̅̅̅̅̅̅̅̅
stage all the input variables that have significant impact – or conversely,
κk = C−kk1 . (11)
have no significant impact – on the output, the effect of each input on the
output variation was ranked and quantified. Due to the quantification, A value of unity for κj indicates true orthogonality of the input f *j to
sensitive variables can be compared by output effect proximity. Non-
all other input terms (main and interaction terms) and values much
sensitive variables can be eliminated from design consideration at this
greater than unity indicate correlated inputs. Finally, R2 is used to
early design stage.
determine the goodness of fit of the ANOVA model:
Due to the possibility of non-monotonic relationships between the
input parameters and the target outputs, the ANOVA method was used; ‖y* − e* ‖2
as was the Sobol’ quasi-random sampling method (Lam et al., 2016; R2 = . (12)
‖y* ‖2
Tian, 2013). Not all input–output relationships will be non-monotonic;
nor will all combinations of inputs in the design space result in non- The Sobol’ sequence is a deterministic algorithm, often referred to as
monotonic output relationships. quasi-random. It provides a uniform sampling of the design space,
The smoothing spline ANOVA (SS-ANOVA) method (Gu, 2013; Ricco reducing the clustering effects of random sampling. The sampled points
et al., 2013) was used in modeFRONTIER. This is a non-parametric ensure maximal distance from each other. It is a low-discrepancy model
regression model that can handle non-linear and non-monotonic re­ – where lower discrepancy means more uniform distribution. All input
lationships. It estimates the mean of an outcome as a smooth function variables are varied at once in this sampling method. It is suitable for
with a decomposition into main and interaction effects where each term non-linear and non-monotonic models (Sobol, 1979).
represents its contribution to the total output variance. Different variables that are commonly considered at the early design
For a given regression problem for an unknown multivariate function stage were tested in the sensitivity analysis. Among the design variables,
the tilt angle is the most significant determinant in solar potential, with

524
S. Yip et al. Solar Energy 224 (2021) 516–530

the optimal average annual solar radiation collected at an angle equal to


0.350
the site latitude (Duffie and Beckman, 2013).
Even though the tilt angle is the most significant determinant in solar 0.300
potential, in practice it may be difficult for BIPV and BIPV/T integrated 0.250

Sensitivity index (-)


projects to attain this design objective – especially in a cold northern
climate where the optimal yearly average solar capture is at latitudes of 0.200
45◦ and higher. For BIPV and BIPV/T roofs, this creates a lot of extra 0.150
interior volume that may not have a practical use (except as attic) but
0.100
which requires a lot more building materials and therefore increases
construction costs. On the exterior, the benefit of the increased roof area 0.050
would be increased deployable BIPV/T area. On the other hand, a lower 0.000
tilt angle may not achieve NZE and have potential excessive snow

Plan shape

WWR S

WWR N

WWR W
Win ID

WWR E

Insul., roof

Insul., wall
accumulation issues in winter, which also result in lower electricity and
useful heat production. Four separate sensitivity analyses were run with
the BIPV/T tilt angle at 30◦ , 35◦ , 40◦ , and 45◦ . Using the Sobol’ method,
κmin = 1.000001 (Insul., wall) Factor
10 000 samples were simulated for each tilt angle. These configurations κmax = 1.00001 (Plan shape) 2
R = 0.845
did not represent any optimization method; that is a topic for further
research. Fig. 6. Main effects sensitivity index for net energy use intensity at 30◦ BIPV/T
roof tilt angle (β).
3. Findings

3.1. Simulation results: insolation, PV yield, EUI 0.350


0.300
The calculated mean daily global insolation for a south-facing sur­
face at a tilt angle equal to the site latitude (equal to 45◦ for Montreal) 0.250
was 4.38 kWh/m2. And the calculated PV yield was approximately Sensitivity index (-) 0.200
1300 kWh/kWp. This compares to a reference of 4.36 kWh/m2 and
0.150
1200 kWh/kWp from Natural Resources Canada’s photovoltaic and
solar resource maps (Natural Resources Canada, 2017). In the simula­ 0.100
tion results, the design configuration with the closest characteristics 0.050
(L15, ψ = − 15◦ , β = 35◦ ) to the Varennes Library yielded an EUI of
63 kWh/m2a without renewables. The most recent published data shows 0.000
ψ

WWR N

WWR W
Plan shape

WWR S

Win ID

WWR E

Insul., roof

Insul., wall
the Varennes Library has an operational EUI of 70 kWh/m2a (Dermar­
diros et al., 2019).
For the purpose of relative energy performance comparisons at the
early design stage, these results are within a reasonable range, within κmin = 1.000001 (Insul., wall) Factor 2
κmax = 1.00001 (Plan shape) R = 0.832
10% or less. A more detailed investigation of the relationship between
the building-integrated PV/T system modeling and the building
morphology is a topic for a future paper. Fig. 7. Main effects sensitivity index for net energy use intensity at 35◦ BIPV/T
roof tilt angle (β).

3.2. Sensitivity analysis: Main effects

0.350
The sensitivity analyses showing main effects are provided in Table 5
and shown in Figs. 6–9 for the four BIPV/T roof tilt angles with respect 0.300
to the target of net annual EUI. The minimum and maximum collinearity 0.250
Sensitivity index (-)

index values, κk , hew closely to unity and the coefficients of determi­


nation, R2, vary between 0.811 and 0.845. 0.200
The plan shape was the most influential for β = 30◦ , and second most 0.150
influential for the other roof tilt angles. In all cases this represented
0.100
between 24.0% and 27.4% of the variance. For β = 35◦ , 40◦ , and 45◦ , the
most influential factor was the building orientation, ψ; its range of effect 0.050
was between 21.4% and 30.5% of the variance. The WWR south was the 0.000
ψ

WWR S

WWR N
Plan shape

Win ID

WWR E

WWR W

Insul., roof

Insul., wall

Table 5
Main effects sensitivity indices for net energy use intensity at four different tilt
angles (β).
κmin = 1.000001 (Insul., wall) Factor 2
Factor β = 30◦ β = 35◦ β = 40◦ β = 45◦ κmax = 1.00001 (WWR W) R = 0.821

Azimuth (ψ) 0.214 0.248 0.279 0.305


Plan shape 0.240 0.246 0.258 0.274 Fig. 8. Main effects sensitivity index for net energy use intensity at 40◦ BIPV/T
WWR south 0.214 0.195 0.174 0.153 roof tilt angle (β).
WWR north 0.166 0.154 0.142 0.130
WinID 0.054 0.050 0.046 0.041 second most influential factor at a tilt angle of 30◦ , and was otherwise
WWR east 0.049 0.047 0.044 0.041
WWR west 0.049 0.046 0.043 0.040
ranked third, representing between 15.3% and 21.4% of the variance.
Insulation, roof 0.008 0.009 0.009 0.010 The only other factor with a sensitivity index over 10% was the WWR
Insulation, wall 0.005 0.005 0.005 0.004 north, with 13.0% to 16.6% depending on the roof tilt angle. The

525
S. Yip et al. Solar Energy 224 (2021) 516–530

22.1%. This result is different from that of the main effects SA. The in­
0.350
clusion of the second-order factors with the pronounced effect of the
0.300 interaction pair of plan shape and building orientation diminished the
0.250
contribution of the plan shape and changed the apportioning of the
Sensitivity index (-)

output variance. For the other β, like in the case of the main effects SA,
0.200 WWR S was the second most influential factor. Its contribution index
0.150 range was between 15.6% and 22.1% – within 3% of results from the
main effects SA.
0.100
The next most significant WWR was that for the north façade, which
0.050 at 12.8% to 16.5% was approximately 18 to 25% less significant, but
0.000 within 2% of results in the main effects. Generally, each of the window
to wall ratios’ influences decrease with increasing BIPV/T tilt angle,
ψ

Plan shape

WWR S

WWR N

WWR E

WWR W
Win ID

Insul., roof

Insul., wall
indicating the increased contribution of the PV electricity generation
over the passive design performance in the net EUI target.
The sensitivity index for plan shape was relatively constant, between
κmin = 1.000001 (Insul., wall) Factor
κmax = 1.00001 (WWR W) 2
R = 0.811
11.3% and 11.6%, across the four tilt angles since the BIPV/T roof base
geometry is defined by the plan shape and this relationship is constant
Fig. 9. Main effects sensitivity index for net energy use intensity at 45◦ BIPV/T regardless of tilt angle. However, its contribution is approximately half
roof tilt angle (β). as large as in the main effects SA. This reduction can be accounted for by
the interaction effect with orientation, as explained further below. The
sensitivity indices for the plan shape, azimuth, WWR south, and WWR sensitivity indices of the other input variables were significantly lower
north were significantly greater than all other factors. The window type than that of the plan shape. In particular the wall and roof insulation
(Win ID), WWR east, and WWR west occupy a less significant plateau of factors were the least sensitive due to their input values being very high,
between 4.0% and 5.4%. Finally, the roof and wall insulation had the compatible with a cold climate.
least significance, between 0.4% and 1.0%. Like in the case of the insulation factors, the window type’s input
values respect the low thermal conductance energy code requirements
for the cold Montreal climate – resulting in a narrow, low sensitivity
3.3. Sensitivity analysis: Interaction effects range of 3.9% to 5.4%. The WWR east and WWR west contribution
indices are within this same range of influence, at 4.0% to 4.9%, and
However, the results are different once interaction effects between 3.8% to 4.7%, respectively. While certain plan shapes (CO, L60, L75,
the factors – where the impact of one factor is dependent on the level of U60, U75) have sizeable east and west façade surface areas when the
another factor – are taken into account. The second order sensitivity building is oriented south, the east and west facades represent – on
indices are provided in Table 6 and shown in Figs. 10–13 for the four average across all simulations and all orientations – effectively the fa­
BIPV/T roof tilt angles with respect to the target of net annual EUI. Each cades with the smallest surface areas. This made their WWR contribu­
interaction pair that accounted for less than 0.005 of the sensitivity tions to the building passive design performance relatively small. As
index was grouped together in the interactions category. Like in the case previously mentioned, the plan shape and orientation input domains
of the main effects SA, the minimum and maximum collinearity index reflect the design choice of solar-enhanced forms with large south- or
values, κk , are close to unity. The coefficients of determination, R2, vary near-south facing areas.
between 0.814 and 0.831. Also, like in the case of the main effects SA, The SA including second-order effects was able to identify an
the results showed the building orientation, ψ, had the most influence on important interaction effect between the orientation and the plan shape
the variance of the output of net annual EUI for β = 35◦ , 40◦ , and 45◦ . that contributed to between 9.1% and 13.6% of the total variance. Since
Quantitatively, it represents between 21.4% and 29.9% of the variance the input domain for the orientation and the geometrical definition of
across all four tilt angles – within 2% of results from the main effects SA. the plan shapes was focused on maximizing solar gains and the output
Within the ψ input domain of − 45◦ to 45◦ , this ranking result was was net EUI, this was an expected result. The path of the sun makes it
anticipated since the planning objective was to create solar-enhanced such that the BIPV/T roof surfaces need to be oriented as best as possible
building forms – hence south- or near-south facing. However, the towards the south to maximize electricity generation. This interaction
contribution of PV electricity generation to the net EUI decreases as the effect was more pronounced at higher tilt angles where there was more
BIPV/T roof slope decreases, thus passive design factors like the WWR electricity generation.
for the south and north facades become more important. This can
explain why at β = 30◦ , the most influential input variable was the 3.4. Effect of BIPV/T roof tilt angle
window to wall ratio for the south façade, with a sensitivity index of
The BIPV/T roof tilt angle has significant impact on the sensitivity of
Table 6 the input variables since the target was net EUI which depends, in part,
Second order sensitivity indices for net energy use intensity at four different tilt on the PV generation. This is most clearly seen in Fig. 14 in which the
angles (β). second-order SA results are compared across all four β angles used. As
Factor β = 30◦ β = 35◦ β = 40◦ β = 45◦ the tilt angle is increased, the combination of increased PV roof area and
Azimuth (ψ) 0.214 0.245 0.274 0.299
favourable inclination results in the PV generation accounting for a
WWR south 0.221 0.199 0.177 0.156 greater portion of the energy balance. The SA input variable which has
WWR north 0.165 0.152 0.140 0.128 influence on the PV generation, the orientation, along with the plan
Plan shape 0.113 0.113 0.114 0.116 shape and orientation interaction pair, increases in significance. On the
Plan shape * Azimuth (ψ) 0.091 0.107 0.121 0.136
other hand, the window to wall ratios for the south and north facades,
WinID 0.054 0.049 0.044 0.039
WWR east 0.049 0.046 0.043 0.040 factors related to passive solar design, decreased in an almost mirror
WWR west 0.047 0.044 0.041 0.038 image rate compared to orientation and the plan shape and orientation
Insulation, roof 0.009 0.009 0.010 0.010 interaction pair, respectively. Put another way, as the tilt angle
Insulation, wall 0.005 0.005 0.005 0.005 increased, the relative significance of the first ranked factor to the other
Interactions 0.031 0.031 0.032 0.032
factors increased as well. At β = 30◦ , the contribution index of the first

526
S. Yip et al. Solar Energy 224 (2021) 516–530

0.350

0.300

0.250

Sensitivity index (-)


0.200

0.150

0.100

0.050

0.000
WWR S

WWR N

Plan shape * ψ

WWR E

WWR W
Plan shape

Win ID

Insul., roof

Insul., wall

Interactions
κmin = 1.001 (Plan shape *ψ) Factor
2
κmax = 1.084 (Insul., roof) R = 0.831

Fig. 10. Second order sensitivity index for net energy use intensity at 30◦ BIPV/T roof tilt angle (β).

0.350

0.300

0.250
Sensitivity index (-)

0.200

0.150

0.100

0.050

0.000
ψ

WWR N
WWR S

Plan shape * ψ

WWR E

WWR W
Plan shape

Win ID

Insul., roof

Insul., wall

Interactions

κmin = 1.001 (Plan shape *ψ) Factor 2


κmax = 1.090 (Insul., roof) R = 0.828

Fig. 11. Second order sensitivity index for net energy use intensity at 35◦ BIPV/T roof tilt angle (β).

ranked factor (WWR south) was 3.2% greater than the second ranked variance is apportioned across all input variables – thus quantifying the
factor (orientation). At β = 45◦ , the contribution index of the first ranked influence of the plan shape with respect to the other input variables on
factor (orientation) was 47.8% greater than the second ranked factor the output variance. This contribution of the plan shape has not been
(WWR south). analyzed in previous papers. As it turns out, from our results, plan shape
Other passive solar design factors like window type, window to wall is important, but not the most influential input variable.
ratio for east and west facades also had a slight decrease in sensitivity Previous SA have not been uniform in their findings for WWR in­
index with increasing β. Of note, the plan shape and the opaque insu­ fluence. At the extremes, WWR was the most influential factor for both
lation variables were practically unaffected by tilt angle. heating and cooling energy performance, separately, (Mechri et al.,
2010); and not very influential for heating energy performance (Garcia
Sanchez et al., 2014). In one case, the WWR combined all windows from
3.5. Discussion all facades and orientations (Mechri et al., 2010), while the other used a
separate WWR for each façade orientation (Garcia Sanchez et al., 2014),
In any global SA, it is important to identify and define the input like in our case. But, unlike either, ours used NZE as the target and found
variables to be analyzed, their appropriate range of values, and the only WWR South and North to be influential. Therefore their analyses
target since these influence the SA results. In this paper, compound plan focused on passive design while ours on both passive and active design.
shapes have been included alongside other, more commonly analyzed, Continuing with these two papers, both found that building
input variables such as WWR. This is important since the output

527
S. Yip et al. Solar Energy 224 (2021) 516–530

0.350

0.300

0.250

Sensitivity index (-)


0.200

0.150

0.100

0.050

0.000
ψ

WWR S

WWR N

WWR E

WWR W
Plan shape * ψ

Plan shape

Win ID

Insul., roof

Insul., wall

Interactions
κmin = 1.001 (Plan shape *ψ) Factor 2
κmax = 1.095 (Insul., roof) R = 0.823

Fig. 12. Second order sensitivity index for net energy use intensity at 40◦ BIPV/T roof tilt angle (β).

0.350

0.300

0.250
Sensitivity index (-)

0.200

0.150

0.100

0.050

0.000
ψ

WWR S

Plan shape * ψ

WWR N

WWR E

WWR W
Plan shape

Win ID

Insul., roof

Insul., wall

Interactions

κmin = 1.001 (Plan shape *ψ) Factor 2


κmax = 1.097 (Insul., roof) R = 0.814

Fig. 13. Second order sensitivity index for net energy use intensity at 45◦ BIPV/T roof tilt angle (β).

orientation was not that influential, though not negligible, for heating Lastly, the two other papers only analyzed rectangular-shaped plans
energy performance (Garcia Sanchez et al., 2014; Mechri et al., 2010) and one concluded that analyzing more complex shapes such as L- or U-
and cooling energy performance (Mechri et al., 2010), but that insu­ shaped plans would be required before being able to extend their results
lation thickness was significant (Garcia Sanchez et al., 2014). In our (Mechri et al., 2010).
case, we found building orientation to be the most influential in three As mentioned in Section 1.3, a global SA can be useful as a screening
out of the four SA, with it being second-most influential in the other tool. At the early design stage, input variables that do not contribute
case; and insulation levels for roof and walls to be the least significant of much to the output variance may be eliminated from consideration. A
all the factors we analyzed. dilemma at early design stage is how to leave the design space as large as
The range the other two papers used for orientation was [0, 180◦ ] possible to avoid overlooking possible design solutions while circum­
whereas we used [− 45◦ , 45◦ ] since the objective was to design a solar scribing the space to limit modeling and computational resources.
enhanced NZEB. Extending the range of the orientation may have led to Since five of the nine input variables analyzed exhibit low sensitivity,
different results but that extended range was viewed as invalid for this the exploration of the design space through an automated simulation
exercise. As for the range of the insulation levels, it was large, but the framework – such as the one described in Section 2.2.2 – using full-
lower end represented an already high performance building meeting factorial design or numerical optimization can be facilitated by
strict energy codes. This may explain why the increment of higher reducing the dimension of the design space by five.
insulation did not produce any sensitivity.

528
S. Yip et al. Solar Energy 224 (2021) 516–530

significant interaction pair of plan shape and orientation.


0.350
Although not an input variable to the SA, the BIPV/T tilt angle had a
significant effect on the SA results. As the tilt angle was increased from
0.300 30◦ to 45◦ , variables associated with PV energy generation, orientation
and the plan shape * orientation interaction pair, increased in signifi­
cance by 40% and 50%, respectively; while variables associated with
ψ
0.250 passive design, WWR S and WWR N, decreased by 30% and 22%
WWR S
respectively. Also, the contribution index of the first ranked factor
Sensitivity index (-)

WWR N
increased with increasing tilt angle. At 45◦ , first ranked orientation was
0.200 Plan shape almost 50% more influential than second ranked WWR south.
Plan shape*ψ Identifying and quantifying orientation, plan shape, window to wall
0.150 Win ID ratios for the south and north facades, and BIPV/T tilt angle as influ­
WWR E ential variables in net zero energy building design are important find­
WWR W ings to assist building designers in focusing on a subset of early design
0.100
Insulation, roof stage variables to reach net zero energy while still leaving the input
Insulation, wall
space large enough for design creativity. The smaller number of input
variables can help reduce the modeling and computational resources
0.050 interactions
necessary in automated design approaches such as full-factorial design
and numerical optimization. Furthermore, in the case of design variable
0.000 restrictions, the remaining significant variables can be prioritized or
30 35 40 45 optimized using the magnitude of their sensitivity as a guide.
Tilt angle, β (°) In the second paper, part 2, the methodology is applied in a case
study for early stage design in which relative comparisons are made
Fig. 14. Comparison of the results of second order sensitivity indices for net considering building form and net-zero energy. This is then extended to
energy use intensity at 30◦ , 35◦ , 40◦ , and 45◦ BIPV/T roof tilt angles (β). include building interactions with the grid: load matching and energy
flexibility. Archetypal design solutions for energy flexibility are
4. Conclusion compared to those for net-zero energy.

This paper, part 1, had as objective to quantify the influence of Declaration of Competing Interest
building form, including courtyards, passive and active design perfor­
mance, and BIPV/T integration on net energy performance in medium- The authors declare that they have no known competing financial
sized institutional NZEBs through sensitivity analysis. To this end, a interests or personal relationships that could have appeared to influence
series of archetypes, including the courtyard, for medium-sized institu­ the work reported in this paper.
tional NZEBs was developed for early design stage analysis – with con­
centration on air-based BIPV/T roof-integrated systems. The Acknowledgements
requirement to reduce unnecessary volume in the attic area limited the
depth of building – which coincided with requirements for enhancing This work was supported by the Natural Sciences and Engineering
daylighting depth for the entire building floor plate. This is a design Research Council of Canada/Hydro-Québec Industrial Research Chair in
constraint specific to northern climates. Higher tilt angles are required Optimized Operation and Energy Efficiency: Towards High Performance
to enhance annual solar capture, usually approximately equal to the site Buildings; and the Fonds de recherche du Québec – Nature et
latitude. technologies.
The simulation workflow based on TRNSYS, Python, and mode­
FRONTIER was presented and a global sensitivity analysis was carried References
out to rank and quantify the most influential input variables commonly
ASHRAE, A., 2016. ASHRAE/IES Standard 90.1-2016: Energy Standard for Buildings
considered at the early design stage using net annual energy use in­ Except Low-Rise Residential Buildings. American Society of Heating, Refrigerating
tensity as the target. Most importantly, for the first time, a sensitivity and Air-Conditioning Engineers, Atlanta.
Athienitis, A., O’Brien, W., 2015. Modeling, Design, and Optimization of Net-Zero
analysis has included the building form as a factor. In all, the combi­
Energy Buildings. Wiley, Hoboken.
nation of medium-sized institutional NZEB archetypes with BIPV/T roof Athienitis, A.K., Bambara, J., O’Neill, B., Faille, J., 2011. A prototype photovoltaic/
in a sensitivity analysis with building form as a factor has previously not thermal system integrated with transpired collector. Sol. Energy 85 (1), 139–153.
been found in the literature. Athienitis, A.K., Barone, G., Buonomano, A., Palombo, A., 2018. Assessing active and
passive effects of façade building integrated photovoltaics/thermal systems:
The most influential input variable across the three steepest BIPV/T Dynamic modelling and simulation. Appl. Energy 209, 355–382.
tilt angles used was the building orientation; the window to wall ratio of Athienitis, A.K., Cellura, M., Chen, Y., Delisle, V., Bourdoukan, P., Kapsis, K., 2015.
the south façade was the most influential for the least steep tilt angle, Modeling and design of Net ZEBs as integrated energy systems. In: Athienitis, A.K.,
O’Brien, W. (Eds.), Modeling, Design, and Optimization of Net-Zero Energy
30◦ . The plan shape, and window to wall ratios for the south and north Buildings. Wiley, Hoboken.
façade were also significant input variables; their sensitivity indices and Buonomano, A., De Luca, G., Montanaro, U., Palombo, A., 2016. Innovative technologies
ranking varied with BIPV/T tilt angle. At the other extreme, all other for NZEBs: An energy and economic analysis tool and a case study of a non-
residential building for the Mediterranean climate. Energy Build. 121, 318–343.
input variables were at least one order of magnitude less sensitive than Canadian Commission on Building and Fire Codes, Construction, I.f.R.i., 2015. National
the most influential variable. Of those, the wall and roof insulation were Energy Code of Canada for Buildings, 2015. National Research Council, Ottawa, ON.
the least sensitive, owing to stringent energy code requirements in a cold Caruso, G., Kämpf, J.H., 2015. Building shape optimisation to reduce air-conditioning
needs using constrained evolutionary algorithms. Sol. Energy 118, 186–196.
climate.
Chen, X., Huang, J., Yang, H., Peng, J., 2019a. Approaching low-energy high-rise
Including second-order factors in the sensitivity analysis was building by integrating passive architectural design with photovoltaic application.
important. This was most clearly evidenced by the different top ranking J. Cleaner Prod. 220, 313–330.
Chen, X., Yang, H., Peng, J., 2019b. Energy optimization of high-rise commercial
result at the shallowest BIPV/T angle (β = 30◦ ) of plan shape when only
buildings integrated with photovoltaic facades in urban context. Energy 1–17.
considering main effects and window to wall ratio for the south façade
when considering interaction effects; and the identification of the

529
S. Yip et al. Solar Energy 224 (2021) 516–530

Chen, X., Yang, H., Zhang, W., 2015. A comprehensive sensitivity study of major passive Natural Resources Canada, 2012. Survey of Commercial and Institutional Energy Use -
design parameters for the public rental housing development in Hong Kong. Energy Buildings 2009. In: Efficiency, O.o.E. (Ed.). Ottawa, Canada.
93, 1804–1818. Natural Resources Canada, 2017. Photovoltaic and solar resource maps. Accessed
Chen, Y., Athienitis, A.K., Galal, K., 2010. Modeling, design and thermal performance of 2019.06.30 2019. https://www.nrcan.gc.ca/18366.
a BIPV/T system thermally coupled with a ventilated concrete slab in a low energy Negendahl, K., 2015. Building performance simulation in the early design stage: an
solar house: Part 1, BIPV/T system and house energy concept. Sol. Energy 84 (11), introduction to integrated dynamic models. Autom. Constr. 54, 39–53.
1892–1907. Nejma, H.B., Guiavarch, A., Lokhat, I., Auzenet, E., Claudon, F., Peuportier, B., 2013. In-
Crawley, D.B., Hand, J.W., Kummert, M., Griffith, B.T., 2008. Contrasting the capabilities situ performance evaluation by simulation of a coupled air source heat pump/PV-T
of building energy performance simulation programs. Build. Environ. 43 (4), collector system. In: IBPSA (Ed.) 13th Conference of International Building
661–673. Performance Simulation Association. IBPSA, Chambéry, France, pp. 1927–1935.
Delisle, V., Kummert, M., 2016. Cost-benefit analysis of integrating BIPV-T air systems Nembrini, J., Samberger, S., Labelle, G., 2014. Parametric scripting for early design
into energy-efficient homes. Sol. Energy 136, 385–400. performance simulation. Energy Build. 68(PART C), 786–798.
Dermardiros, V., Athienitis, A.K., Bucking, S., 2019. Energy performance, comfort, and O’Brien, W., Bourdoukan, P., Delisle, V., Yip, S., 2015. Net ZEB design processes and
lessons learned from an institutional building designed for net zero energy. ASHRAE tools. In: Athienitis, A.K., O’Brien, W. (Eds.), Modeling, Design, and Optimization of
Trans. 125 (Part 1), 682–695. Net-Zero Energy Buildings. Wiley, Hoboken, pp. 107–174.
Duffie, J.A., Beckman, W.A., 2013. Solar engineering of thermal processes, fourth ed. Ourghi, R., Al-Anzi, A., Krarti, M., 2007. A simplified analysis method to predict the
Wiley, Hoboken. impact of shape on annual energy use for office buildings. Energy Convers. Manage.
ESTECO SpA, 2017. modeFRONTIER 2017R2, 2017R2 ed. ESTECO SpA, Trieste, Italy. 48 (1), 300–305.
Garcia Sanchez, D., Lacarrière, B., Musy, M., Bourges, B., 2014. Application of sensitivity Paulson, B.C., 1976. Designing to reduce construction costs. J. Constr. Div. 102 (4),
analysis in building energy simulations: combining first- and second-order 587–592.
elementary effects methods. Energy Build. 68(PART C), 741–750. Pessenlehner, W., Mahdavi, A., 2003. Building morphology, transparence, and energy
Government of Canada, 2019. 1981-2010 Climate Normals and Averages, in: Canada, E. performance. In: Eighth International IBPSA Conference 2003. IBPSA, Eindhoven,
a.C.C. (Ed.). Government of Canada, Ottawa. NL, pp. 1025–1032.
Granadeiro, V., Duarte, J.P., Correia, J.R., Leal, V.M.S., 2013. Building envelope shape Peterson, K., Torcellini, P., Grant, R., 2015. A Common Definition for Zero Energy
design in early stages of the design process: Integrating architectural design systems Buildings. US Department of Energy, Washington, DC, p. 22.
and energy simulation. Autom. Constr. 32, 196–209. Python Software Foundation, 2019. Python, 3.7.3 ed. Python Software Foundation,
Gu, C., 2013. Smoothing Spline ANOVA Models, second ed. Springer, New York, NY. Beaverton, OR.
Guglielmetti, R., Pless, S., Torcellini, P., 2010. On the use of integrated daylighting and Ratti, C., Raydan, D., Steemers, K., 2003. Building form and environmental performance:
energy simulations to drive the design of a large net-zero energy office building. In: Archetypes, analysis and an arid climate. Energy Build. 35 (1), 49–59.
Proc. Fourth National Conference of IBPSA-USA, New York, NY. Ricco, L., Rigoni, E., Turco, A., 2013. Smoothing spline ANOVA for variable screening,
Hachem, C., Athienitis, A., Fazio, P., 2011. Parametric investigation of geometric form Dolomites Research Notes on Approximation. Trento, Italy, pp. 130–139.
effects on solar potential of housing units. Sol. Energy 85 (9), 1864–1877. Rounis, E.D., Athienitis, A.K., Stathopoulos, T., 2016. Multiple-inlet Building Integrated
Heiselberg, P., Brohus, H., Hesselholt, A., Rasmussen, H., Seinre, E., Thomas, S., 2009. Photovoltaic/Thermal system modelling under varying wind and temperature
Application of sensitivity analysis in design of sustainable buildings. Renew. Energy conditions. Sol. Energy 139, 157–170.
34 (9), 2030–2036. Sartori, I., Napolitano, A., Voss, K., 2012. Net zero energy buildings: a consistent
Hemsath, T.L., Alagheband Bandhosseini, K., 2015. Sensitivity analysis evaluating basic definition framework. Energy Build. 48, 220–232.
building geometry’s effect on energy use. Renew. Energy 76, 526–538. Sherman, M.H., 1987. Estimation of infiltration from leakage and climate indicators.
Hopfe, C.J., Hensen, J.L.M., 2011. Uncertainty analysis in building performance Energy Build. 10 (1), 81–86.
simulation for design support. Energy Build. 43 (10), 2798–2805. Sobol, I.M., 1979. On the Systematic Search in a Hypercube. SIAM J. Numer. Anal. 16
Kamel, R.S., Fung, A.S., 2014. Modeling, simulation and feasibility analysis of residential (5), 790–793.
BIPV/T+ASHP system in cold climate - Canada. Energy Build. 82, 758–770. Straube, J.F., Burnett, E.F.P., 2005. Building Science for Building Enclosures. Building
Kämpf, J.H., Montavon, M., Bunyesc, J., Bolliger, R., Robinson, D., 2010. Optimisation of Science Press, Westford, Mass.
buildings’ solar irradiation availability. Sol. Energy 84 (4), 596–603. Tian, W., 2013. A review of sensitivity analysis methods in building energy analysis.
Kämpf, J.H., Robinson, D., 2010. Optimisation of building form for solar energy Renew. Sustain. Energy Rev. 20, 411–419.
utilisation using constrained evolutionary algorithms. Energy Build. 42 (6), Torcellini, P., Pless, S., Deru, M., Crawley, D., 2006. Zero Energy Buildings: A Critical
807–814. Look at the Definition, ACEEE Summer Study. U.S. Department of Energy, Pacific
Kazanci, O.B., Skrupskelis, M., Sevela, P., Pavlov, G.K., Olesen, B.W., 2014. Sustainable Grove, California, p. 12.
heating, cooling and ventilation of a plus-energy house via photovoltaic/thermal Tuhus-Dubrow, D., Krarti, M., 2010. Genetic-algorithm based approach to optimize
panels. Energy Build. 83, 122–129. building envelope design for residential buildings. Build. Environ. 45 (7),
Mitchell, J.C., Braun, J.E., Evans, B.L., 2017. TRNSYS 18: A Transient System Simulation 1574–1581.
Program, 18 ed. Solar Energy Laboratory, University of Wisconsin, Madison, Voss, K., Musall, E., 2011. Net Zero Energy Buildings: International Comparison of
Wisconsin. Carbon-Neutrality in Buildings. Birkhäuser Verlag, Basel.
Konis, K., Gamas, A., Kensek, K., 2016. Passive performance and building form: An Wang, W., Rivard, H., Zmeureanu, R., 2006. Floor shape optimization for green building
optimization framework for early-stage design support. Sol. Energy 125, 161–179. design. Adv. Eng. Inf. 20 (4), 363–378.
Lam, T.C., Ge, H., Fazio, P., 2016. Energy positive curtain wall configurations for a cold Yang, T., Athienitis, A.K., 2015. Experimental investigation of a two-inlet air-based
climate using the Analysis of Variance (ANOVA) approach. Build. Simul. 9 (3), building integrated photovoltaic/thermal (BIPV/T) system. Appl. Energy 159,
297–310. 70–79.
Lawrence Berkeley National Laboratory, 2016. WINDOW, v7.4 ed. Lawrence Berkeley Yi, Y.K., Malkawi, A.M., 2009. Optimizing building form for energy performance based
National Laboratory, Berkeley. on hierarchical geometry relation. Autom. Constr. 18 (6), 825–833.
Li, H., Wang, S., Cheung, H., 2018. Sensitivity analysis of design parameters and optimal Yip, S., Chen, Y., Athienitis, A., 2015. Comparative Analysis of a Passive and Active
design for zero/low energy buildings in subtropical regions. Appl. Energy 228, Daylight Redirecting Blind in Support of Early Stage Design, CISBAT 2015. ÉPFL,
1280–1291. Lausanne, Switzerland, pp. 229–234.
Marszal, A.J., Heiselberg, P., Bourrelle, J.S., Musall, E., Voss, K., Sartori, I., Youssef, A.M.A., Zhai, Z.J., Reffat, R.M., 2016. Genetic algorithm based optimization for
Napolitano, A., 2011. Zero Energy Building - A review of definitions and calculation photovoltaics integrated building envelope. Energy Build. 127, 627–636.
methodologies. Energy Build. 43 (4), 971–979. Zhang, L., Zhang, L., Wang, Y., 2016. Shape optimization of free-form buildings based on
Martin, L., March, L., 1972. Urban Space and Structures. Cambridge University Press, solar radiation gain and space efficiency using a multi-objective genetic algorithm in
London. the severe cold zones of China. Sol. Energy 132, 38–50.
Mechri, H.E., Capozzoli, A., Corrado, V., 2010. USE of the ANOVA approach for sensitive
building energy design. Appl. Energy 87 (10), 3073–3083.

530

You might also like