Next Article in Journal
Does Rheology of Bingham Fluid Influence Upscaling of Flow through Tight Porous Media?
Next Article in Special Issue
Utilising Open Geospatial Data to Refine Weather Variables for Building Energy Performance Evaluation—Incident Solar Radiation and Wind-Driven Infiltration Modelling
Previous Article in Journal
Protection Method Based on Wavelet Entropy for MMC-HVDC Overhead Transmission Lines
Previous Article in Special Issue
The Use of Energy Models in Local Heating Transition Decision Making: Insights from Ten Municipalities in The Netherlands
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Energy Modelling and Analytics in the Built Environment—A Review of Their Role for Energy Transitions in the Construction Sector

by
Massimiliano Manfren
1,
Maurizio Sibilla
2,* and
Lamberto Tronchin
3
1
Faculty of Engineering and Physical Sciences, University of Southampton, Boldrewood Innovation Campus, Burgess Rd, Southampton SO16 7QF, UK
2
School of the Built Environment, Oxford Brookes University, Headington Campus, Oxford OX3 0BE, UK
3
Department of Architecture (DA), University of Bologna, Viale Europa 596, 47521 Cesena, Italy
*
Author to whom correspondence should be addressed.
Energies 2021, 14(3), 679; https://doi.org/10.3390/en14030679
Submission received: 15 December 2020 / Revised: 20 January 2021 / Accepted: 25 January 2021 / Published: 28 January 2021
(This article belongs to the Special Issue Open Data and Models for Energy and Environment)

Abstract

:
Decarbonisation and efficiency goals set as a response to global warming issue require appropriate decision-making strategies to promote an effective and timely change in energy systems. Conceptualization of change is a relevant part of energy transitions research today, which aims at enabling radical shifts compatible with societal functions and market mechanisms. In this framework, construction sector can play a relevant role because of its energy and environmental impact. There is, however, the need to move from general instances to specific actions. Open data and open science, digitalization and building data interoperability, together with innovative business models could represent enabling factors to accelerate the process of change. For this reason, built environment research has to address the co-evolution of technologies and human behaviour and the analytical methods used for this purpose should be empirically grounded, transparent, scalable and consistent across different temporal/spatial scales of analysis. These features could potentially enable the emergence of “ecosystems” of applications that, in turn, could translate into projects, products and services for energy transitions in the built environment, proposing innovative business models that can stimulate market competitiveness. For these reasons, in this paper we organize our analysis according to three levels, from general concepts to specific issues. In the first level, we consider the role of building energy modelling at multiple scales. In the second level, we focus on harmonization of methods for energy performance analysis. Finally, in the third level, we consider emerging concepts such as energy flexibility and occupant-centric energy modelling, considering their relation to monitoring systems and automation. The goal of this research is to evaluate the current state of the art and identify key concepts that can encourage further research, addressing both human and technological factors that influence energy performance of buildings.

Graphical Abstract

1. Introduction

In recent years, a notable research effort has been devoted to the conceptualisation of sustainability transitions [1] and, more specifically for energy, to the identification of “complementarities” at multiple levels [2,3]. Transition processes embody the necessity of radical-shifts and they represent an opportunity for innovation and entrepreneurship [4], with a clear focus on issues such as global warming and decarbonisation of energy systems [5]. In these innovation processes, the role of intermediaries and strategic niches appears to be crucial. In fact, understanding how actors can control and accelerate the energy transition is a key issue for research today [6] and intermediaries can play a fundamental role in this direction [7]. Intermediaries (i.e., public, non-profit, and private third-parties [8]) are actors which facilitate relations between key actors and enable knowledge sharing and pooling [9].The opportunities for the construction industry in this sense are relevant, because of the impact of built environment in terms of raw resources, energy and carbon emissions [10], but also because of the potential to exploit innovative technologies within emerging paradigms such as circular economy [11]. There is, however, the need to move from general instances to specific actions. These actions have to enable radical shifts compatible with societal functions and market mechanisms; for this reason, in this research we focus on energy modelling and analytics that can provide critical insights in this sense. At present, it possible to identify multiple enabling factors for radical shifts and acceleration of the process of change. First, the evolution of practices focused on concepts such as open data, open innovation, open science [12,13,14] and, in particular, open energy modelling principles [15,16]. Second, advances in building data interoperability (technical, informational and organizational) [17] and data availability at multiple levels, using technologies such as the Internet of Things (IoT) [18,19,20] and cyber-physical systems [21], which can enable, in turn, innovation in end-user energy delivery [22], and in energy infrastructures [23]. Third, the increasing decentralization of energy systems where the co-evolution of built environment and energy infrastructures [24] plays a fundamental role, that can be investigated by means of “soft-linking” of energy modelling approaches, from planning to operation [25]. Finally, innovative business models proposing concepts such as prosumer [26] and prosumager [27], which are determining changes in the way energy market works and energy trading takes place, for example using Peer-to-Peer automated exchange mechanisms, exploiting Blockchain technologies [28].
In this rapidly evolving framework, research aimed at radical changes in energy systems and built environment needs to consider the enabling factors reported above and to acknowledge the limitations and bottlenecks in view of energy efficiency and carbon reduction goals. The aim of this paper is to discuss to what extent and in what ways energy modelling and analytics can support the process of change for energy transitions in the construction sector. In Section 2 we illustrate the background of the research, explaining the fundamental elements that motivate it.

2. Background and Motivation

Energy transitions involve the transformation of the network of players and organisations traditionally working in the energy sector (e.g., policy-makers, regulators, transmission and distribution authorities, etc.) as well as the change of the role of customers, from passive to active (i.e., prosumers [26] and prosumagers [27]).In fact, socio-technical innovations are critically dependent on the possibility to access new information, knowledge and resources, which are key enablers for the development of innovative products and services [29], within a market mechanism. Construction sector can be conceptualized, for example, by considering three fundamental domains [30]: project, product and service. All these domains are going to be deeply influenced by socio-technical changes in energy transitions, which will transform the way buildings are designed, built and managed. Sharing knowledge among actors is crucial when addressing building energy performance in a comprehensive way, considering both human and technical factors [31]. In fact, the impact of occupants has to be considered from multiple stand-points [32] and users’ behaviour can determine both “re-bound” [33] and “pre-bound” effects [34,35], that can create a substantial difference between expected and measured performance, which can be inscribed in the general category of “performance gaps” [36,37,38]. A “performance gap” can be found in all the stages of building life cycle [39] and the use of standardized assumption in modelling, e.g., to create Energy Performance Certificates, has to be critically questioned when using them to estimate actual energy consumption and potential savings [40].
Additionally, the dynamic interaction between building and energy infrastructures [41,42] has to be considered as well for multiple reasons (e.g., operational constraints, limitations of the penetration of renewables, innovative business model for the electricity market, etc.) and in light of possible developments in terms of “soft-linking” of energy models [25]. Finally, considering building performance from a whole life cycle perspective (indeed critical for emerging paradigms such as circular economy [11]), embodied energy in materials, technologies and processes represents another potential “performance gap” to be considered [43,44]. In fact, all these potential gaps create risks and lack of credibility when investing in energy efficiency and sustainability measures. Therefore, monitoring, verifying and tracking performance (i.e., energy, emission and cost in particular) using robust, transparent and empirically grounded methods is essential to evaluate the effectiveness of measures and share knowledge regarding practices. This, in turn, can contribute to investment de-risking and stimulate the growth of business “ecosystems” in energy and sustainability transitions, particularly for the construction sector. Additionally, the co-benefits of energy efficiency measures (e.g., improved indoor environmental quality, health, productivity, pollution reduction, etc.) [45] have to be considered both by policy makers and investors, to weight properly cost and benefits. Following the general trend towards open science, briefly outlined in Section 1, the research community in the energy field has stressed in recent years the fundamental importance of open energy data and models [46,47] and we can envisage an evolution towards systems of model [48] designed to address key problems in energy transitions, eventually taking advantage of “soft-linking” approaches [25,49]. Rather than being designed for separate applications, models can be potentially conceived and work like “ecosystems” [48] of interconnected applications, based on open data and modelling standard [46] where the researchers are opening their modelling “black-boxes” [47]. Indeed, transparent and robust models can become part of innovative business strategies, leading to techno-economically feasible pathways in transitions(thereby enabling a radical change to happen in practice). In fact, this review is part of a more extensive research work focused on “Buildings-as-Energy-Service” concept, in which separate literature reviews were conducted to explore both social and physical science perspectives on this topic. The concepts emerging from the reviews represent the basic elements of a Cognitive Mapping [50] process. The aim of this process is to create an inter-disciplinary research environment (a cognitive framework) [51] that is essential for innovation processes, where creativity is stimulated by the participation of user in the process of knowledge creation and sharing [52]. In Section 3 we describe the research methodology used to identify the role of energy modelling and analytical techniques in relation to the issues mentioned above.

3. Research Methodology

Considering the issues briefly outlined in Section 1 and Section 2, the objective of this review study is to identify and analyse the features of energy modelling and analytical techniques that could be enabling factors in energy transition processes. The two fundamental research questions posed in this study are the following. First, what are the modelling techniques that can meet the criteria that will be described later in this section? Second, what are the essential characteristics (of modelling approaches) that can contribute to reduce the level of fragmentation of knowledge? The modelling framework proposed as outcome of the research attempts to reduce the level of fragmentation of the highly diversified body of knowledge available and to help in the conceptualization of processes of change (energy transition) by identifying opportunities, together with limitations and bottlenecks.
In this research both qualitative and quantitative data are analysed and it is therefore a “mixed approach” [53]. For this reason, we used concepts from Grounded Theory [54] as a reference for our research, in which both qualitative and quantitative data are utilised (“all is data” [55]). In brief, Grounded Theory (GT) can be defined as a “a set of integrated conceptual hypotheses systematically generated to produce an inductive theory about a substantive area” [56] and as “theory that was derived from data, systematically gathered and analysed through the research process” [57]. The results of a GT study are “a set of concepts, related to each other in an interrelated whole” [58].
The limitations of such approach depend on the fact that the selection in literature sampling depend on the subjective judgment (point of view) of the researcher and cannot stand outside of it [58]. However, the process can become more transparent and reproducible by stating the steps and the criteria used in it. In this research, we followed seven steps:
(1)
Definition of knowledge domains of interest;
(2)
Stratified search using domain and keywords in Web of Science database (WoS);
(3)
Initial selection of pertinent literature on WoS;
(4)
Definition of additional criteria for inclusion/exclusion of literature;
(5)
Initial verification of literature using title, keywords and abstract;
(6)
Final selection of literature;
(7)
Detailed analysis of literature.
The fundamental knowledge domain of interest is “Building Energy Performance” (step 1) and the keywords considered initially are “Building stock”, “Uncertainty” and “Flexibility” (step 2), to address fundamental topics in research. “Building stock” is chosen to identify examples of building energy modelling at multiple scales (e.g., for planning and policy, utility scale studies, etc.). “Uncertainty” is chosen to identify studies that analyse the critical dimension of energy performance uncertainty, which may create risks and lack of credibility for efficiency practices, starting from fundamental principles in Measurement and Verification (M&V) and Monitoring & Targeting (M&T). “Flexibility” is chosen to identify research regarding the interaction between building and infrastructures, which is strictly related to their technological co-evolution. The results obtained in step 2 are summarized in Table 1.
In order to obtain the final literature selection, additional criteria have been introduced and re-sampling of literature has been conducted iteratively until “theoretical saturation” was reached. Theoretical saturation term indicates “the phase of qualitative data analysis in which the researcher has continued sampling and analysing data until no new data appear and all concepts of the theory are well-developed and their linkages to other concepts are clearly described” [59].The criteria used in re-sampling have been summarized and motivated in Table 2. They are derived from previous research in the area of energy modelling [24,60] and consider the general trends towards the use of open data for energy research [46] and the necessity to increase of transparency in energy modelling [47]. In other words, the criteria introduced represent, in our opinion, limiting factors and constraints for the creation of “ecosystems” of models [48], which are briefly outlined in Section 2.
In Section 4 the results of the review process are presented, structuring them according to three levels of analysis (related to the domain and keyword chosen, as explained before in this section) that correspond to the development, by means of iterative sampling, of the key concepts reported in Table 1. The overall research process is synthesized graphically in Figure 1.

4. Results and Discussion

In this section we discuss how energy modelling and analytical tools could support energy transition processes for the construction industry, highlighting relevant insights for research across the three levels of analysis introduced in Section 3. The three levels proposed are indeed a strategy to perform a decomposition of the problem, going from general principles to specific issues that are emerging within the research framework. In Section 4.1 we analyse the topic of building energy performance analysis at multiple scales and its implications (e.g., in energy and planning policy, utility scale studies, etc.), which introduces the issues at general level (first level of analysis). In Section 4.2 we present harmonized methodologies (based on M&V principles and considering possible extensions) to analyse energy performance in buildings and we synthesize their characteristics (second level of analysis). Finally, in Section 4.3, we introduce innovative topics such as energy flexibility (infrastructures’ interaction) and occupant-centric (users’ interaction) energy modelling, which will contribute to redefine how buildings are actually designed and operated in the future (third level of analysis). Overall, throughout these three levels we show how many of the ongoing research developments are deeply related to the fundamental elements that motivate our research and are described in Section 2.

4.1. Building Energy Performance Analysis at Multiple Scales

Comprehensive reviews of building energy models have been published in recent years [61,62,63] and, while energy performance is particularly relevant, more comprehensive approaches to building performance analysis [64] are crucial for the evolution of the building sector. As anticipated, the analysis of building energy performance requires an understanding of both human and technical factors [31], and this confirms the inherent socio-technical dimension of energy modelling and analytics. It is therefore necessary to structure energy performance analysis with respect to both human and technical factors. In turn, this is important, for example, to address properly the gap between design and measured performance, i.e., the performance gap [36,37,38], introduced in Section 2. Further, the concept of statistical “Reference Buildings” [65] (RB) must be introduced to enable building performance benchmarking at multiple scales. RB models represent the common typologies, technologies and end-uses in the building stock, identified through statistical analysis and expert knowledge (e.g., on building technologies, types of end-uses, user behaviour, etc.) on a large-scale base. Building data are usually multi-level data, which makes it difficult to access the full information needed to describe in detail the performance of building stock. However, building energy modelling data can be organised in a hierarchical and standardized way; examples in this sense can be found at the EU level in the legislation on the definition of cost-optimal performance levels [66] and in EU Building Stock Observatory [67]. Further, in the US, technical standardisation has been tested with the definition of RB models [68,69], accounting also for the costs of various technological options [70].The role of energy modelling cycles and the importance of the level of detail (from conceptual to final design) are considered by the standard ASHRAE 209 [71]. Additionally, the use of hierarchical structures in datasets for building energy modelling can be found, for example, in performance gap studies [37], in the analysis of impact of automation systems [72], and in occupancy modelling [73].Further, with respect to building energy model calibration on measured data, we can find examples using multi-level data [74] and exploiting macro-parameters [75] (i.e., lumped quantities) to facilitate and guide the uncertainty and sensitivity analysis, together with the use of archetypes [76] (i.e., RB for a certain construction typology), and of additional information such as monitored internal temperature profiles [77].At the state of the art, multiple modelling options are available, depending on the scope of the analysis process, which range from physics based (“law driven”) “white-box” models to statistics and machine leaning based (“data driven”) “black-box” models. An analysis of the suitability of the different modelling strategies has been proposed by Koulamas et al. [78] and, more specifically for model calibration, by Manfren et al. [79]. Indeed, it is possible to use models to simulate performance (forward modelling) and to estimate model inputs from measured performance (inverse modelling) in multiple ways. Therefore, using forward and inverse modelling techniques [24] in a synergic way for calibration purposes is crucial. In this context, advanced techniques such as Bayesian analysis can help reconstructing built stock data under uncertainty [80,81,82], using probabilistic ranges for the model input parameters. The possibility to benchmark building performance on a large scale base [83,84] can increase the effectiveness of policies and can guarantee better decision-making processes, not only for policy makers but for multiple stakeholders (e.g., designers, energy managers, investors, etc.). In fact, the progressive convergence of bottom-up and top-down perspectives in energy modelling and planning for building stock [61] can contribute to the development of “soft-linking” approaches between various types of models [25] and, consequently, ensure consistency of actions in transition processes at multiple levels. Overall, a systematic statistical approach to building performance analysis [85] can be crucial to the evolution of design and operation paradigms for building stock. In recent years we assisted to an increasing commitment towards energy efficiency in buildings which led to the definition of paradigms such as Passive House [86], NZEB [87,88], and PEB [89], considering just the most relevant. Indeed, the possibility to deploy these paradigms at scale is subject to technical and economic constraints. In this sense, the use of statistical “Reference Buildings” can support techno-economic optimization studies [65,90], utility scale analysis of design [91] and operation of buildings [92] and energy planning at national scale [68,69,70], where innovative building paradigms are proposed and implemented. In terms of computation, the necessity of performing parametric (or probabilistic) simulation studies [93,94,95] is emerging and the algorithmic definition of simplified building models [96,97,98] can be exploited for building stock modelling at city scale [99,100,101] and regional scale [102]. In Table 3 we synthesize the outcomes of literature analysis regarding building energy performance analysis at multiple scales, highlight the main target of the different studies and their scale of analysis, namely national, regional, urban and stock. The latter indicates, in general, studies that are proposing building performance analysis on multiple typologies and end-uses.
The examples reported before are clearly not exhaustive but they are used to illustrate the potential role of building energy performance analysis at large scale, using modelling methods that are transparent and reproducible, build upon (or compatible with) technical standardization. These topics are developed further in Section 4.2, consider two fundamental dimensions: the quantification of the impact of energy efficiency measures and the ability model dynamic behaviour (i.e., load profiles). Finally, at the beginning of this Section we stressed the importance of a precise hierarchy for multi-level building energy modelling data. Another important aspect is that of “vertical integration” of information in energy modelling, from user up to infrastructures (e.g., user, individual spaces within the room, individual rooms, building zones, whole building, meter, energy infrastructure). Examples of research in this direction can be found in IEA Annexes on “Energy Flexibility in Buildings” [105] and “Occupant-Centric Building Design and Operation” [106]. These fundamental aspects of current research are discussed more in detail in Section 4.3.

4.2. Harmonizing Methodologies to Analyse Energy Performance

Appropriate spatial and temporal resolution of data is necessary to track building energy performance at multiple scales and energy metering data constitute, of course, the basic information layer. There is the need for harmonized methods that can ensure robust evidence (empirically grounded and validated) for efficiency measures (not only for research, but also for policy), by means of reliable statistics regarding the actual impact of efficient technologies [107,108] and especially by means of performance benchmarking of efficiency measures [109,110]. The term “harmonized” is used here to indicate, in general, methodologies in which redundancies and overlapping features are removed; harmonized methods can help documenting performance transparently, for example by tracking evidence of energy efficiency savings (and also related carbon and cost savings) in time and detecting the impact of influencing factors. Measurement and Verification (M&V) protocols [111,112] and methods represent the backbone in this sense and important research initiatives have been conducted in recent years to enhance and extend their applicability, such as the Uniform Methods Project (UMP) and other related projects [109,110,113]. The goal of these projects was harmonising the methods for the quantification of energy savings for different efficiency measures, both in residential and commercial buildings. Multiple measures (technologies) are included (HVAC, HP/chillers, CHP, lighting, envelope, variable-frequency drives, etc.). Another important project, focused on de-risking investment in energy efficiency, is the Investor Confidence Project (ICP) [114]. As already mentioned, the methods used in these projects represent an extension of the ones that can be found in M&V protocols [111,112] and technical standards [115,116,117], in which thresholds (expressed as statistical KPIs, representing the “goodness of fit”) are given for the acceptability of models as calibrated [118] on measured data. Finally, open software is available [113,119,120] as a basis for further development that can potentially be enabled by open science principles (i.e., transparency and reproducibility of results, among others).
In general, these approaches are based on energy interval data (dependent variable) and weather data (independent variables) along with other independent variables (e.g., dummy variables for models of various occupancy and operational regimes) which can be derived from contextual knowledge and information. Instead of using energy data directly, it is possible to use the energy signature [115], which is the average power over the number of hours of operation in the interval considered. The most important independent variable for weather normalization of energy consumption is outdoor air temperature [121,122] and these methods are affine to variable-base degree days methods [92,123]. Temperature response methods are reviewed by Fazeli et al. [124]. Conceptual simplicity is one of their advantages (among others), compared to other meta-modelling techniques [125,126]. Automated model selection techniques [119,127] can be applied as well to compare the performance of multiple modelling options, using statistical KPIs representing their “goodness of fit”. From an analytical perspective, it is important to be able to connect both the design and the operation phase analysis [128,129] in order to ensure consistency in the use of energy performance analysis techniques over the different phases of the life cycle [130]. In this way reliable limits for performance measured or estimated [131] can be produced and used against benchmarks, allowing a continuous improvement process (i.e., Plan Do Check Act is one of the key principles of Energy Management Systems [132]).
Far from being merely instruments for weather normalisation of energy use (i.e., outdoor temperature dependence), harmonised approaches can also help modelling dynamic loads (e.g., demand response) [109], ideally clustering operating conditions for typical profiles [133,134,135] to obtain specific insights on recurrent operating schedules (e.g., depending on the type of end-use).
In reality, understanding load dynamics at multiple scales is crucial for providing accurate estimates of the impact of flexibility measures that can inform policy [136] by creating a “soft link” between modelling approaches. Load modelling techniques can be used to complement “traditional” optimization approaches in cases where they are no longer sufficient and several operational configurations need to be studied [137]. Furthermore, the possibility of evaluating the thermal, electrical and fuel requirements with harmonised methods can extend further the principle of “soft-linking” of energy models in multi-commodity systems [138,139,140,141,142]. In this sense, harmonised methods should complement (in terms of general principles) open science-based approaches to energy research [16] because of their transparency. In addition, they may help to address related issues such as energy demand forecasts in future climate change scenarios [143,144,145] and definition of load profiles evolution due to efficiency measures and behavioural change, which are fundamental for optimizing decentralised energy systems in buildings [146] and communities [138,147,148].
In short, harmonised approaches can be used to discuss two main aspects of energy modelling research in a rigorous and transparent manner: the quantification of the effect of energy efficiency measures and the reconstruction of dynamic behaviour (i.e., time series modelling), such as load profiles analysis. Table 4 below provides a comparison of the main features of regression-based modelling methods that can meet the constraints set out in Section 3. We consider different types of end-uses, namely residential and non-residential, and different types of energy services, namely heating, cooling, domestic hot water (DHW), and appliances. First of all, the selected and reviewed literature reflects, in large part, empirically based studies in which the authors used operation phase data. The research is performed in all cases using regression-based (interpretable) methods that are significantly consistent with the harmonisation and standardisation principles outlined in this section. In terms of temporal scalability, the papers are categorised with respect to monthly, daily and hourly data. In certain cases, sub-hourly data are used, but we classify them as hourly data since this is the highest resolution considered by the model calibration thresholds proposed in the standards and protocols [118]; in any case, this resolution is adequate to capture the essence of building dynamic energy behaviour. In terms of spatial scalability, we consider building subsystems (building fabric and technological systems), building as a whole, building stock, and community and city scale. For the latter, the term design corresponds substantially to planning; the operational phase data are used as a basis for making accurate forecasts for the future. In addition, whole building energy balance is used in most situations, although in some cases (e.g., evaluation of building fabric characteristics) the energy balance at the zone or room level is used. Finally, with the term approximate physical approximation, we suggest the possibility of using regression coefficients to estimate physical quantities. Overall, the table illustrates how harmonized/standardized regression-based methods can cover several temporal and spatial scales of analysis and how they can theoretically combine design and operational phase performance analysis into the same analytical workflow (thereby satisfying re-configurability criteria, reported in Table 2). Finally, regression models can be used for both residential and non-residential end-uses to study energy services (heating, cooling, DHW, appliances) in multiple ways and can provide insights up to building system level when sub-metering data (e.g., thermal, electric) are available, while enabling, at the same time, the aggregation of results on a large scale base for building stock modelling.
The possibility to employ advanced harmonized analytical techniques could, in principles, contribute to the development of innovative business models built upon Energy Performance Contracting (EPC) [179] principles, where dynamic operational conditions are clustered [134] and multiple regression models are combined together [156] to investigate performance, integrating data at multiple spatial and temporal resolutions, while retaining an approximated physical interpretation. Further, the graphical representation of regression-based methods can be combined with other visualization strategies used for energy (and exergy) flows at multiple scales, from building systems and sub-systems [180], to networks in multi-energy systems [181]. Physical-statistical (i.e., “grey-box”) formulations [158,173,182,183,184,185], can extend the inherent capabilities of these modelling approaches even further and provide additional insights that may be particularly valuable in a continuous improvement logic, while retaining scalability [183,184].
Despite the variety of possible model formulations, we believe that data-driven approaches should use energy modelling definitions and quantities that are consistent with those proposed in the current technical standardization [186] to improve the comparability of results and consistency with policy objectives, for which standardisation plays a key role. For this reason, we report hereafter in Table 5 some experimental protocols (harmonized or standardized) with examples of applications at component level and building zone level. Indeed, the table highlights the potential continuity and integration of these experimental methods to estimate thermo-physical properties of building components and zones. Ideally, they could partially overlap with methods presented in Table 4, for example by alternating short-term measurement at higher frequency with long-term measurement at lower frequency [157] during building life cycle.
In QUB and ISABELE methods, the definitions used are in line with current technical standardisation; the physical parameters are represented by lumped quantities (thus reducing the number of parameters needed) and the model formulation greatly reduces the complexity compared to a physical “white-box” model, briefly recalled in Section 4.1. “White-box” models are detailed models based on physical laws used mainly for simulations during the design process and validated in accordance with energy simulation test standards [194,195].The potential contact point between “white-box” detailed modelling and “grey-box” (physical-statistical) lumped modelling parameters can be found in multi-level building energy model calibration [74] where “macro-parameters” (aggregated, lumped quantities) [75] are used to validate more detailed models, together with additional information such as internal temperature profiles [77] and other contextual information.
Indeed, the potential advantages of “grey-box” models are that they can be derived (and verified) from the basic concepts of energy analysis [196,197], built by using highly standardised rules [188], and they can employ efficient state-space [198] and analytical formulations [199]. Examples of validation of “grey-box” models using simulation test standards at the state of the art have been published by Lundström et al. [195] and Michalak [194,200]; a “grey-box” model for the detection of thermo-physical properties by inverse modelling has been implemented also in EnergyPlus, a detailed “white-box” modelling software [201]. Juricic et al. [202] considered the effect of natural weather variability in the identification of building envelope characteristics using these model types, showing how approximately two weeks of data are sufficient to achieve adequate accuracy. Finally, Baasch et al. [203] compared the performance of different “grey-box” methods in the derivation of thermo-physical properties from smart thermostat data acquisition (i.e., directly from temperature data instead of energy and temperature data), showing promising results.
“Grey-box” models can be also converted to “black-box” (i.e., statistical and machine learning models) for specific applications, for example control [204] or monitoring of internal conditions [205,206].“Black box” models are computationally efficient but they need to be trained on data before being deployed. As a result, “grey-box” models can be viewed as an intermediate stage between “white-box” and “black-box” models, and many examples of implementations have been found in recent years, ranging from experimental test facilities for building technologies [207] and construction components [208], to incorporation into the Building Information Modeling (BIM) workflow [209], and even to integrated room automation [210].
In addition, regression-based and “grey-box” model capabilities can be used in the Bayesian analysis framework. Bayesian analysis is suitable, for example, to ‘reconstruct’ building data (by estimating its characteristics) under uncertainty [80,81,82] or to evaluate the robustness of “grey-box” model estimates with respect to variable operating conditions [211] using Monte Carlo simulation methods [212], to reproduce realistically uncertain operating conditions.
What appears to be important for future research in this area is to increase the transparency of the modelling process by means of harmonised methodologies (using uniform rules and interpretable models as shown above) in order to verify and monitor output efficiently and to boost their level of automation without increasing complexity unnecessarily. Furthermore, the role of building automation [72,213] and monitoring systems [214,215] is crucial to understand the real dynamic behaviour of buildings by means of detailed data that can of course, complement energy metering, which represents the basic level of knowledge. Surrogate physical-statistical models (i.e., “grey-box” models) can be implemented also as “digital twins” (i.e., digital reproductions of the dynamic behaviour of their physical counterparts) at the level of construction technologies [216,217]. As a conclusion, in this Section we highlighted how harmonized methods for energy performance analysis are essential from multiple stand-points and how statistical and physical-statistical approaches are crucial for the evolution of energy research in buildings. Indeed, the methods reported and discussed in this Section can complement research on energy demand in end-uses based on epidemiology concept [218,219], providing however robust evidence on the performance of technologies and systems using empirically grounded methods, based on M&V principles.

4.3. Energy Flexibility and Occupant-Centric Energy Modelling

Energy flexibility in buildings [105] and occupant-centric energy modelling [106] for building design and operation are important research topics at present and they are directly addressing changes in fundamentals components of energy systems, such as users and energy infrastructures. Therefore, the topics discussed in this Section are complementing the ones in Section 4.1, focused on the potential of building performance analysis at scale, and Section 4.2, focused on harmonised methods for energy performance analysis (static and dynamic), showing how innovative concepts can contribute to reshape building design and operation strategies in the future. The analysis of the “mismatch” between building load profiles and on-site generation profiles (e.g., using PV power generation) has received a great deal of attention in recent years [41], due to the necessity of managing electric grid with increasing penetration of renewables. In this context, the concept of energy flexibility has been introduced to account for the dynamic interaction between end-user and electric infrastructures. Energy flexibility can be defined as the ability to control demand and supply according to consumer needs, grid conditions and climate [220]; an extensive review on this concept has been written by Reynders et al. [42]. There exist multiple options for increasing flexibility at the energy system level [136] and “soft-linking” of modelling approaches is increasingly important for energy planning and operation purpose [25,137]. More specifically, flexibility in buildings depends on the ability to use storage resources and to act on devices (including HVAC) after a trigger (e.g., time, power, energy price, etc.). Heating Ventilation and Air Conditioning (HVAC)systems are crucial because of their impact on the overall consumption of buildings and because of the potentially active role in energy infrastructure for demand response [221] and for absorbing surplus of energy from renewables [222]. From a technical perspective, energy flexibility in buildings can be exploited to shape building load profiles or to maximize the amount of energy that is self-consumed on-site [223,224], thereby increasing the matching between demand and on-site generation. The flexibility potential can be determined by the thermal inertia of building construction components (thermal mass) and by the presence of technical systems with storage (thermal and/or electric). Indeed, the exploitation of on-site renewables in buildings requires the adoption of technologies such as photovoltaics, heat pumps and energy storage [225]. Further, on the infrastructure side, flexibility requires an evolution of standardization of communication protocols to ensure efficient operation [226] and the results in this sense can determine a relevant change for the electric energy system as a whole [227], which may be combined with (and pushed forward by) consumer centric innovations in business models [228]. Specific KPIs [229] are required to describe flexibility potential and a large part of research at the state of the art concentrates on strategies to unlock it by means of control strategies [229,230], considering also related topics such as appropriate levels of modelling complexity and effort for their implementation [231]. In Table 6 we report an analysis of control strategies aimed at building flexibility for different end-uses and services using the same abbreviations as in Table 4. In Table 6 we consider the control objective in relation to flexibility, namely Load Shaping (LS) and On-site Renewable Maximization (ORM), following the arguments reported above. Additionally, the control types considered are Rule-Base Control (RBC), Optimal Control (OC) and Model Predictive Control (MPC). In Rule-Base Control rules are designed to fulfil a certain control objective but are not designed to achieve optimization of the overall system behaviour. In Optimal Control the control strategy is defined as an objective function to be optimized but doesn’t include a prediction for the future. In Model Predictive Control the strategy is defined by means of an optimization performed with a certain control horizon (usually 24/48h); a comprehensive review on MPC has been written by Drgona et al. [232]. Further, we indicate the technical elements on which control strategies are focused. Also in this case, control strategies can be used for both residential and non-residential buildings a can exploit flexibility of heating, cooling and DHW demand by using the thermal storage capabilities of building fabric and technical system (e.g., water storage tanks). What appears to be fundamental, both in predictive and non-predictive cases, is the definition of dynamic operating schedules and set-points trajectories that are constrained by comfort requirements for heating and cooling services. However, the implementation of a detailed comfort model is challenging, due to the characteristics of control-oriented modelling approaches, and, for this reason, simplifications are generally considered when defining operational boundaries (i.e., the constraints for operation). Finally, the dynamic interaction with the grid is particularly important when dynamic tariffs are present and optimized control strategies have to consider the cost of imported and exported energy on a dynamic base.
It is worth noticing that there exists a potential methodological continuity between M&V practices at the state of the art, presented in Section 4.2, and innovative control strategies that represent an evolution of weather compensated control. This can be achieved, for example, using dynamic re-setting of heating and cooling curves [234] and machine learning algorithms whose performance can be tested and compared transparently in different weather conditions [250]. In general, by integrating regression modelling and clustering, it is possible to analyse variations of dynamic operational trajectories [134,156]. User behaviour has a huge impact on all the building services reported in Table 3 and, in recent years, an increasing research effort has been put on “Occupant-Centric Building Design and Operation” [106], as already mentioned before in the text.In particular, extensive reviews on this broad topic have been published recently [32], describing tools, methods and applications; more specific reviews have been dedicated to occupancy and behaviour modelling [254] and to occupant-centric control strategies [255]. The practical necessity to adapt modelling strategies in response to the purpose of the specific study (e.g., design, management, etc.) is indicated with the term “fit-for-purpose” [73]. Considering energy performance in a whole life cycle perspective, the variability of people behaviour and occupancy patterns has to be considered already at the early design stage, in particular in high efficiency and Nearly Zero Energy Buildings (NZEBs) [256]. After that, in the operation stage, occupancy can be measured in different ways [257] and data can be used to conduct realistic simulations [258]. In any case, as reported before, modelling occupancy patterns and user behaviour may require strategies that are customized (i.e., “fit-for-purpose”) for the specific problem to be addressed: one possible solution is that of generating parametric or probabilistic occupancy profiles and modelling all the related variables (e.g., internal gains due to people and appliances, air change rates, etc.) in a transparent way [259,260]. This approach has been used, for example, to analyse building performance gap [261]. Realistic occupancy profiles are fundamental to address not only energy services but also to investigate related issues such as thermal comfort [262], Indoor Environmental Quality (IEQ) [263,264,265] and electric load profiles [266], among others.
As a conclusion, what appears to be important for future research in this area is increasing the transparency of the modelling process and linking it to harmonized methodologies (presented in Section 4.2) to verify and track performance efficiently without increasing unnecessarily the complexity of models themselves (i.e., maintaining an appropriate balance). Further, the role of building automation and monitoring systems is critical to understand the real dynamic behaviour of buildings. For example, data collected by monitoring systems [214,215] and/or automation systems [72,213] enable the performance characterization of envelope [160] and technical systems [267], together with occupancy patterns [257], already mentioned. Building performance monitoring and modelling can exploit also advances in IoT technologies [268] and open software [269], leading to innovative applications for energy and environmental management [270]. The possibility to rely on a combination of simulation methods and empirically grounded techniques for M&V can open interesting research opportunities in these areas.

4.4. Summary of Research Findings

In this section we describe the concepts emerging from studies that are in the intersections of the three levels of analysis presented in Section 4.1, Section 4.2 and Section 4.3, respectively. For this reason, we report in Table 7 the source, the level of analysis and the relevant concepts for the integration of energy modelling and data analytical processes. First, we can see how statistical reference buildings and parametric modelling represent the necessary basis for building energy modelling at multiple scales [80,81,82,103]. After that, “white-box” and “grey-box” modelling approaches can be integrated using a hierarchical multi-level approach [74] where “macro-parameters” [75] (aggregated, lumped quantities) are used as a mean to validated/calibrate more detailed model [80,118]. In turn, “grey-box” models based on regression and time series can guarantee empirically grounded “boundaries” for the estimation of building performance (providing harmonized methods) that may be used in multiple applications, while retaining a physical interpretation of the coefficients. The interpretability of models can provide multiple insights that can be exploited for the continuous improvement of technologies and practices (i.e., the PDCA approach [132]). Additionally, by combining regression, time series and clustering [134,156] it could possible to identify recurrent patterns in user behaviour [73,106] and in infrastructures’ interaction [25,105,136], with a more precise quantification of the actual flexibility achievable. Both aspects (user behaviour and infrastructures’ interaction) have to be considered in innovative business models for buildings where traditional Energy Performance Contracting is combined with innovative features [179] to ensure competitiveness and adequate level of services. Finally, data from automation and monitoring systems [72,160,213,214,215] are necessary to enable in depth analysis of performance, even though dynamic energy metering can be considered as the fundamental layer of information [214,215].
As explained above, energy modelling and data analytical processes can be integrated in systems of models. Ideally, the creation of systems of standardized or harmonized “surrogate” physical-statistical models (i.e., “grey-box” models), which can be implemented in cyber-physical systems could represent a major breakthrough for energy modelling research. It can guarantee, for example, the possibility to act coherently at multiple levels in energy systems, using data analytics as a common background, and to create a certain degree continuity of performance analysis process during building life cycle, from design to operation phase. As discussed in Section 4.2, this result may be achieved by means of regression-based modelling approaches that combine conceptual simplicity and ease of implementation with adequate performance, in terms of analytics. In the next Section with indicate future research work that can be based on the outcomes of this research.

5. Further Work

Further research work could focus on knowledge mapping to enhance the integration and transparency of data within a modelling framework for energy in buildings, able to act at multiple levels. In Section 4.4 we described the points of contact between the multiple levels of analysis considered and we indicated how “surrogate” physical-statistical models (i.e., “grey-box” models that can be implemented in cyber-physical systems) could potentially work in “ecosystems” of applications. “Ecosystems” of models can address different types of end-uses (i.e., residential and non-residential), technological domains (i.e., heating, cooling, DHW, appliances) and applications (e.g., energy management, control, fault detection, environmental monitoring, etc.) while sharing a set of common underlying principles and rules. In this sense, surrogate models can act as “digital twins,” that is to say digital reproductions of the dynamic behaviour of their physical counterparts (or systems). Harmonization and technical standardization play an essential role to avoid redundancy, multiplication of efforts and unnecessary increase of complexity of procedures. In fact, this could be the case of technical issues affecting multiple levels of information in the built environment, such as energy efficiency and flexibility or behavioural modelling and occupant-centric design and operation, described in Section 4.3. As mentioned in the introduction, building data interoperability [17] using common data exchange formats is necessary to increase the digitalisation and automation of buildings. The use of semantic web technologies [271] and standards based on IFC could support not only design but also operation (e.g., energy and environmental monitoring) [272], employing “surrogate” modelling strategies (physical/statistical, “grey-box”) [209] compatible with the above mentioned principles. Finally, as introduced in Section 2, the research presented in this paper is part of a broader investigation, focused on the concept of “Buildings-as-Energy-Service”: new forms of knowledge integration are needed to develop innovative services and products that can work as “ecosystems” and exploit this concept.

6. Conclusions

Energy transitions involve the transformation of the network of players and organisations that have traditionally worked in the energy sector along with new roles for customers. Radical innovation in the energy sector will have an impact on multiple domains in the construction sector (e.g., project, product and service). In this paper, we reviewed ongoing research on energy modelling and analytical tools that could support energy transition processes for the construction sector. In particular, we discussed how harmonised methods for analysing and tracking energy performance (Section 4.2) and innovative concepts such as flexibility and occupant-centric design and operation (Section 4.3) could contribute to a radical change in the built environment, using similar principles of analysis for actions that involve multiple scales (Section 4.1).The review process has been articulated according to three levels of analysis, introduced in Section 3 and reported in Section 4, ranging from general concepts to specific issues and we provided a summary of research findings as a set of interrelated concepts (Section 4.4). Overall, we identified criteria for energy modelling and analytical techniques (i.e., empirically grounding, scalability, harmonization, interpretability and re-configurability), that, in our opinion, constitute constraints to the creation of “ecosystems” of energy models aimed at supporting energy transition processes at multiple levels in the built environment. Regarding the first level of analysis (Section 4.1), systems of models can contribute to the creation of robust empirically grounded studies regarding efficiency for energy policy and utility scale actions. With respect to the second level (Section 4.2), they can be used to integrate data at multiple temporal and spatial scales, streamlining the analytical workflow (starting from consolidated M&V and M&T practices) and they can provide approximated physical interpretation of results, thereby increasing the transparency of modelling. Finally, in the third level (Section 4.3) they can help increasing energy flexibility in the interaction with infrastructures and improving the level of energy services in an occupant centric (design and operation) perspective. In all the levels considered in this review, we stressed the importance of studies that are empirically grounded and that can provide robust evidence for informing future research and policy.
As discussed in Section 5, these principles can constitute the basis for further research work, focused on developing specific applications built on top of them. In fact, the research proposed is part of a broader research activity focused on the “Buildings-as-Energy-Service” concept and the creation of a Tool Kit for knowledge integration regarding this topic, with the support of Cognitive Mapping technique. New forms of knowledge integration are needed to develop innovative services and products and this Tool Kit may be used to engage multiple users in the process of knowledge creation and sharing. Conceptualization is fundamental in innovation studies for energy and sustainability transitions but while general concepts can be clearly understood, what is still unclear is how these concepts can then translate into specific projects, products and services for energy transitions in the built environment, using innovative business models. Tools for knowledge integration can give a contribution in this sense.
Further, the problem of data accessibility has to be considered as well. The lack of detailed data or inadequate data reliability due to non-standardized collection procedures can be addressed using harmonized methodologies (described in Section 4.2). At present, this is causing a knowledge gap that undermines informed policy choices in the energy transition process (as well as in many other processes). Sensors, the Internet of Things (IoT), together with processes of automation and digitalisation described in this paper, could enable access to a greater amount of data for the building stock. In this context, it will be important to create open data repositories about technology, energy demand for end uses and weather data. Standardized and up-to-date data could enable transparent and consistent modelling processes at multiple scales of analysis, partially reducing the effort and stimulating the development of innovative energy technologies and services.
As a conclusion, in this paper we proposed a reflection on concepts that can help structuring future R&D activities and we highlighted a potential way to increase transparency, robustness and reproducibility in modelling by linking general principles emerging from the state of the art of research, to specific applications, employing harmonized methods as the core element. We believe that sharing information and making it more transparent and easily accessible can support multiple communities involved in R&Dfor energy transitions overcoming social and technical issues that may hinder the radical shifts that are necessary for long-term built environment sustainability.

Author Contributions

Conceptualization, M.M., M.S. and L.T.; Methodology, M.M. and M.S.; Investigation, M.M. and M.S.; Writing—original draft preparation, M.M.; Writing review and editing, M.M. and L.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by UK Research and Innovation through the Industrial Strategy Challenge Fund, Transforming Construction Network Plus subcontract CID 3835572.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The project “Developing a Tool Kit for Knowledge Integration: Envisioning Buildings-as-Energy-Service” is supported by The Transforming Construction Network Plus which is funded by UK Research and Innovation through the Industrial Strategy Challenge Fund. The N+ unites construction’s academic and industrial communities to create a new research and knowledge base, dedicated to addressing the systemic problems holding back the sector. The N+ is a joint project between UCL, Imperial College London and WMG, University of Warwick.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Köhler, J.; Geels, F.W.; Kern, F.; Markard, J.; Onsongo, E.; Wieczorek, A.; Alkemade, F.; Avelino, F.; Bergek, A.; Boons, F.; et al. An agenda for sustainability transitions research: State of the art and future directions. Environ. Innov. Soc. Transit. 2019, 31, 1–32. [Google Scholar] [CrossRef] [Green Version]
  2. Markard, J.; Hoffmann, V.H. Analysis of complementarities: Framework and examples from the energy transition. Technol. Forecast. Soc. Chang. 2016, 111, 63–75. [Google Scholar] [CrossRef]
  3. Sibilla, M.; Kurul, E. Distributed Renewable and Interactive Energy Systems in Urban Environments. TECHNE J. Technol. Archit. Environ. 2018, 1, 33–39. [Google Scholar] [CrossRef]
  4. Carayannis, E.G.; Campbell, D.F.J. Mode 3 Knowledge Production in Quadruple Helix Innovation Systems: 21st-Century Democracy, Innovation, and Entrepreneurship for Development; Springer: Berlin/Heidelberg, Germany, 2011; ISBN 9781461420620. [Google Scholar]
  5. Carayannis, E.G.; Barth, T.D.; Campbell, D.F. The Quintuple Helix innovation model: Global warming as a challenge and driver for innovation. J. Innov. Entrep. 2012, 1, 2. [Google Scholar] [CrossRef] [Green Version]
  6. Gliedt, T.; Hoicka, C.E.; Jackson, N. Innovation intermediaries accelerating environmental sustainability transitions. J. Clean. Prod. 2018, 174, 1247–1261. [Google Scholar] [CrossRef]
  7. Smith, A.; Raven, R.R. What is protective space? Reconsidering niches in transitions to sustainability. Res. Policy 2012, 41, 1025–1036. [Google Scholar] [CrossRef] [Green Version]
  8. Kolk, A.; Van Tulder, R.; Kostwinder, E. Business and partnerships for development. Eur. Manag. J. 2008, 26, 262–273. [Google Scholar] [CrossRef] [Green Version]
  9. Bush, R.E.; Bale, C.S.E.; Powell, M.; Gouldson, A.; Taylor, P.G.; Gale, W.F. The role of intermediaries in low carbon transitions—Empowering innovations to unlock district heating in the UK. J. Clean. Prod. 2017, 148, 137–147. [Google Scholar] [CrossRef]
  10. Berardi, U. A cross-country comparison of the building energy consumptions and their trends. Resour. Conserv. Recycl. 2017, 123, 230–241. [Google Scholar] [CrossRef]
  11. Barrie, J.; Zawdie, G.; João, E. Leveraging triple helix and system intermediaries to enhance effectiveness of protected spaces and strategic niche management for transitioning to circular economy. Int. J. Technol. Manag. Sustain. Dev. 2017, 16, 25–47. [Google Scholar] [CrossRef]
  12. Open Science Policy Platform (OSPP). Available online: https://ec.europa.eu/research/openscience/index.cfm?pg=open-science-policy-platform (accessed on 25 August 2020).
  13. Mendez, E.; Lawrence, R.; MacCallum, C.J.; Moar, E.; Lossau, N.; Deketelaere, K.; Luyben, K.; Epure, M.; Bertero, M.; Garfinkel, M.; et al. Progress on Open Science: Towards a Shared Research Knowledge System. Final Report of the Open Science Policy Platform. 2020. Available online: https://ec.europa.eu/research/openscience/pdf/ec_rtd_ospp-final-report.pdf (accessed on 5 December 2020).
  14. European Open Science Cloud (EOSC) of the European Commission. Available online: https://ec.europa.eu/research/openscience/index.cfm?pg=open-science-cloud (accessed on 25 August 2020).
  15. Openmod Open Energy Modelling Initiative (Openmod)—Open Models. Available online: https://wiki.openmod-initiative.org/wiki/Open_Models (accessed on 25 August 2020).
  16. Hilpert, S.; Kaldemeyer, C.; Krien, U.; Günther, S.; Wingenbach, C.; Plessmann, G. The Open Energy Modelling Framework (oemof)—A new approach to facilitate open science in energy system modelling. Energy Strateg. Rev. 2018, 22, 16–25. [Google Scholar] [CrossRef] [Green Version]
  17. Hardin, D.; Stephan, E.G.; Wang, W.; Corbin, C.D.; Widergren, S.E. Buildings Interoperability Landscape; United State Department of Energy: Oak Ridge, TN, USA, 2015. [CrossRef] [Green Version]
  18. Atzori, L.; Iera, A.; Morabito, G. The Internet of Things: A survey. Comput. Netw. 2010, 54, 2787–2805. [Google Scholar] [CrossRef]
  19. Tan, L.; Wang, N. Future internet: The Internet of Things. In Proceedings of the 2010 3rd International Conference on Advanced Computer Theory and Engineering (ICACTE), Chengdu, China, 20–22 August 2010; Volume 5, pp. 5–376. [Google Scholar]
  20. Breiner, S.; Subrahmanian, E.; Sriram, R.D. Modeling the Internet of Things: A Foundational Approach. In Proceedings of the Seventh International Workshop on the Web of Things, Stuttgart, Germany, 7 November 2016; ACM: New York, NY, USA, 2016; pp. 38–41. [Google Scholar]
  21. Schmidt, M.; Åhlund, C. Smart buildings as Cyber-Physical Systems: Data-driven predictive control strategies for energy efficiency. Renew. Sustain. Energy Rev. 2018, 90, 742–756. [Google Scholar] [CrossRef] [Green Version]
  22. Reka, S.S.; Dragicevic, T. Future effectual role of energy delivery: A comprehensive review of Internet of Things and smart grid. Renew. Sustain. Energy Rev. 2018, 91, 90–108. [Google Scholar] [CrossRef]
  23. Arghandeh, R.; Von Meier, A.; Mehrmanesh, L.; Mili, L. On the definition of cyber-physical resilience in power systems. Renew. Sustain. Energy Rev. 2016, 58, 1060–1069. [Google Scholar] [CrossRef] [Green Version]
  24. Tronchin, L.; Manfren, M.; Nastasi, B. Energy efficiency, demand side management and energy storage technologies—A critical analysis of possible paths of integration in the built environment. Renew. Sustain. Energy Rev. 2018, 95, 341–353. [Google Scholar] [CrossRef]
  25. Dominković, D.F.; Junker, R.G.; Lindberg, K.; Madsen, H. Implementing flexibility into energy planning models: Soft-linking of a high-level energy planning model and a short-term operational model. Appl. Energy 2020, 260, 114292. [Google Scholar] [CrossRef]
  26. Zafar, R.; Asif, M.; Razzaq, S.; Ali, W.; Naeem, U.; Shehzad, K. Prosumer based energy management and sharing in smart grid. Renew. Sustain. Energy Rev. 2018, 82, 1675–1684. [Google Scholar] [CrossRef]
  27. Sioshansi, F.P. Consumer, Prosumer, Prosumager: How Service Innovations Will Disrupt the Utility Business Model; Academic Press: Cambridge, MA, USA, 2019. [Google Scholar]
  28. Andoni, M.; Robu, V.; Flynn, D.; Abram, S.; Geach, D.; Jenkins, D.P.; McCallum, P.; Peacock, A. Blockchain technology in the energy sector: A systematic review of challenges and opportunities. Renew. Sustain. Energy Rev. 2019, 100, 143–174. [Google Scholar] [CrossRef]
  29. Gui, E.M.; MacGill, I. Typology of future clean energy communities: An exploratory structure, opportunities, and challenges. Energy Res. Soc. Sci. 2018, 35, 94–107. [Google Scholar] [CrossRef]
  30. Thuesen, C.; Koch-Ørvad, N.; Maslesa, E. Organising Sustainable Transition: Understanding the Product, Project and Service Domain of the Built Environment. In Proceedings of the 32nd Annual ARCOM Conference, Manchester, UK, 5–7 September 2016. [Google Scholar]
  31. Yoshino, H.; Hong, T.; Nord, N. IEA EBC annex 53: Total energy use in buildings—Analysis and evaluation methods. Energy Build. 2017, 152, 124–136. [Google Scholar] [CrossRef] [Green Version]
  32. Azar, E.; O’Brien, W.; Carlucci, S.; Hong, T.; Sonta, A.; Kim, J.; Andargie, M.S.; Abuimara, T.; El Asmar, M.; Jain, R.K.; et al. Simulation-aided occupant-centric building design: A critical review of tools, methods, and applications. Energy Build. 2020, 224, 110292. [Google Scholar] [CrossRef]
  33. Herring, H.; Roy, R. Technological innovation, energy efficient design and the rebound effect. Technovation 2007, 27, 194–203. [Google Scholar] [CrossRef] [Green Version]
  34. Sunikka-Blank, M.; Galvin, R. Introducing the prebound effect: The gap between performance and actual energy consumption. Build. Res. Inf. 2012, 40, 260–273. [Google Scholar] [CrossRef]
  35. Rosenow, J.; Galvin, R. Evaluating the evaluations: Evidence from energy efficiency programmes in Germany and the UK. Energy Build. 2013, 62, 450–458. [Google Scholar] [CrossRef]
  36. de Wilde, P. The gap between predicted and measured energy performance of buildings: A framework for investigation. Autom. Constr. 2014, 41, 40–49. [Google Scholar] [CrossRef]
  37. Imam, S.; Coley, D.A.; Walker, I. The building performance gap: Are modellers literate? Build. Serv. Eng. Res. Technol. 2017, 38, 351–375. [Google Scholar] [CrossRef] [Green Version]
  38. de Wilde, P. The building performance gap: Are modellers literate? Build. Serv. Eng. Res. Technol. 2017, 38, 757–759. [Google Scholar] [CrossRef] [Green Version]
  39. Van Dronkelaar, C.; Dowson, M.; Spataru, C.; Mumovic, D. A Review of the Regulatory Energy Performance Gap and Its Underlying Causes in Non-domestic Buildings. Front. Mech. Eng. 2016, 1, 17. [Google Scholar] [CrossRef] [Green Version]
  40. Cozza, S.; Chambers, J.; Deb, C.; Scartezzini, J.-L.; Schlüter, A.; Patel, M.K. Do energy performance certificates allow reliable predictions of actual energy consumption and savings? Learning from the Swiss national database. Energy Build. 2020, 224, 110235. [Google Scholar] [CrossRef]
  41. Salom, J.; Marszal, A.J.; Widén, J.; Candanedo, J.; Lindberg, K.B. Analysis of load match and grid interaction indicators in net zero energy buildings with simulated and monitored data. Appl. Energy 2014, 136, 119–131. [Google Scholar] [CrossRef]
  42. Reynders, G.; Lopes, R.A.; Marszal-Pomianowska, A.; Aelenei, D.; Martins, J.; Saelens, D. Energy flexible buildings: An evaluation of definitions and quantification methodologies applied to thermal storage. Energy Build. 2018, 166, 372–390. [Google Scholar] [CrossRef]
  43. Pomponi, F.; Moncaster, A. Scrutinising embodied carbon in buildings: The next performance gap made manifest. Renew. Sustain. Energy Rev. 2018, 81, 2431–2442. [Google Scholar] [CrossRef] [Green Version]
  44. De Wolf, C.; Pomponi, F.; Moncaster, A. Measuring embodied carbon dioxide equivalent of buildings: A review and critique of current industry practice. Energy Build. 2017, 140, 68–80. [Google Scholar] [CrossRef]
  45. International Energy Agency. Capturing the Multiple Benefits of Energy Efficiency; IEA: Paris, France, 2014. [Google Scholar]
  46. Pfenninger, S.; Decarolis, J.; Hirth, L.; Quoilin, S.; Staffell, I. The importance of open data and software: Is energy research lagging behind? Energy Policy 2017, 101, 211–215. [Google Scholar] [CrossRef]
  47. Pfenninger, S.; Hirth, L.; Schlecht, I.; Schmid, E.; Wiese, F.; Brown, T.; Davis, C.; Gidden, M.J.; Heinrichs, H.; Heuberger, C.; et al. Opening the black box of energy modelling: Strategies and lessons learned. Energy Strateg. Rev. 2018, 19, 63–71. [Google Scholar] [CrossRef]
  48. Bollinger, L.; Davis, C.; Evins, R.; Chappin, E.; Nikolic, I. Multi-model ecologies for shaping future energy systems: Design patterns and development paths. Renew. Sustain. Energy Rev. 2018, 82, 3441–3451. [Google Scholar] [CrossRef]
  49. Deane, J.P.; Chiodi, A.; Gargiulo, M.; Gallachóir, B.P.Ó. Soft-linking of a power systems model to an energy systems model. Energy 2012, 42, 303–312. [Google Scholar] [CrossRef]
  50. Novak, J.D. A Theory of Education: Meaningful Learning Underlies the Constructive Integration of Thinking, Feeling, and Acting Leading To Empowerment for Commitment and Responsibility. Mean. Learn. Rev. 2011, 1, 1–14. [Google Scholar]
  51. Sibilla, M. A meaningful mapping approach for the complex design. Int. J. Des. Sci. Technol. 2017, 23. Available online: http://ijdst.europia.org/index.php/ijdst/article/view/2 (accessed on 10 December 2020).
  52. Florida, R. Cities and the Creative Class. City Community 2003, 2, 3–19. [Google Scholar] [CrossRef]
  53. Creswell, J.W.; Creswell, J.D. Research Design: Qualitative, Quantitative, and Mixed Methods Approaches; SAGE Publications: Thousand Oaks, CA, USA, 2017. [Google Scholar]
  54. Birks, M.; Mills, J. Grounded Theory: A Practical Guide; SAGE Publications: Thousand Oaks, CA, USA, 2015. [Google Scholar]
  55. Glaser, B.G. Doing Grounded Theory: Issues and Discussions; Sociology Press: Mill Valley, CA, USA, 1998. [Google Scholar]
  56. Glaser, B.G.; Holton, J. Remodeling grounded theory. Forum Qual. Soc. Res. 2004, 5, 4. [Google Scholar]
  57. Corbin, J.; Strauss, A. Basics of Qualitative Research: Techniques and Procedures for Developing Grounded Theory, 3rd ed.; SAGE Publications: Thousand Oaks, CA, USA, 2008; ISBN 9781412906432. [Google Scholar]
  58. Charmaz, K. Constructionism and the grounded theory method. Handb. Constr. Res. 2008, 1, 397–412. [Google Scholar]
  59. Lewis-Beck, M.; Bryman, A.; Liao, T.F. The SAGE Encyclopedia of Social Science Research Methods; College of Liberal Arts and Sciences: Chicago, IL, USA, 2004. [Google Scholar]
  60. Manfren, M.; Nastasi, B.; Groppi, D.; Garcia, D.A. Open data and energy analytics—An analysis of essential information for energy system planning, design and operation. Energy 2020, 213, 118803. [Google Scholar] [CrossRef]
  61. Kavgic, M.; Mavrogianni, A.; Mumovic, D.; Summerfield, A.; Stevanovic, Z.; Djurovic-Petrovic, M. A review of bottom-up building stock models for energy consumption in the residential sector. Build. Environ. 2010, 45, 1683–1697. [Google Scholar] [CrossRef]
  62. Foucquier, A.; Robert, S.; Suard, F.; Stéphan, L.; Jay, A. State of the art in building modelling and energy performances prediction: A review. Renew. Sustain. Energy Rev. 2013, 23, 272–288. [Google Scholar] [CrossRef] [Green Version]
  63. Fumo, N. A review on the basics of building energy estimation. Renew. Sustain. Energy Rev. 2014, 31, 53–60. [Google Scholar] [CrossRef]
  64. de Wilde, P. Building Performance Analysis; Wiley & Sons: Hoboken, NJ, USA, 2018; ISBN 9781119341932. [Google Scholar]
  65. Corgnati, S.P.; Fabrizio, E.; Filippi, M.; Monetti, V. Reference buildings for cost optimal analysis: Method of definition and application. Appl. Energy 2013, 102, 983–993. [Google Scholar] [CrossRef]
  66. European Commission. Commission Delegated Regulation; (EU) No 244/2012; European Commision: Brussels, Belgium, 2012. [Google Scholar]
  67. Arcipowska, A.; Rapf, O.; Faber, M.; Fabbri, M.; Tigchelaar, C.; Boermans, T.; Surmeli-Anac, N.; Pollier, K.; Dal, F.; Sebi, C.; et al. Support for Setting up an Observatory of the Building Stock and Related Policies; Buildings Performance Institute Europe (BPIE): Berlin, Germany, 2016. [Google Scholar]
  68. Deru, M.; Field, K.; Studer, D.; Benne, K.; Griffith, B.; Torcellini, P.; Liu, B.; Halverson, M.; Winiarski, D.; Rosenberg, M.; et al. U.S. Department of Energy Commercial Reference Building Models of the National Building Stock; NREL: Denver, CO, USA, 2011.
  69. Goel, S.; Athalye, R.A.; Wang, W. Enhancements to ASHRAE Standard 90.1 Prototype Building Models; No. PNNL-23269; Pacific Northwest National Laboratory (PNNL): Richland, WA, USA, 2014. [Google Scholar]
  70. Thornton, B.A.; Rosenberg, M.I.; Richman, E.E.; Wang, W.; Xie, Y.; Zhang, J.; Cho, H.; Mendon, V.V.; Athalye, R.A.; Liu, B. Achieving the 30% Goal: Energy and Cost Savings Analysis of ASHRAE Standard 90.1-2010; No. PNNL-20405; Pacific Northwest National Laboratory (PNNL): Richland, WA, USA, 2011. [Google Scholar]
  71. ASHRAE. 209-2018—Energy Simulation Aided Design for Buildings Except Low-Rise Residential Buildings; (ANSI Approved); ASHRAE: Washington, DC, USA, 2018. [Google Scholar]
  72. Aste, N.; Manfren, M.; Marenzi, G. Building Automation and Control Systems and performance optimization: A framework for analysis. Renew. Sustain. Energy Rev. 2017, 75, 313–330. [Google Scholar] [CrossRef]
  73. Gaetani, I.I.; Hoes, P.P.-J.; Hensen, J.L. Occupant behavior in building energy simulation: Towards a fit-for-purpose modeling strategy. Energy Build. 2016, 121, 188–204. [Google Scholar] [CrossRef]
  74. Yang, Z.; Becerik-Gerber, B. A model calibration framework for simultaneous multi-level building energy simulation. Appl. Energy 2015, 149, 415–431. [Google Scholar] [CrossRef]
  75. Rodríguez, G.C.; Andrés, A.C.; Muñoz, F.D.; López, J.M.C.; Zhang, Y. Uncertainties and sensitivity analysis in building energy simulation using macroparameters. Energy Build. 2013, 67, 79–87. [Google Scholar] [CrossRef]
  76. Korolija, I.; Marjanovic-Halburd, L.; Zhang, Y.; Hanby, V.I. UK office buildings archetypal model as methodological approach in development of regression models for predicting building energy consumption from heating and cooling demands. Energy Build. 2013, 60, 152–162. [Google Scholar] [CrossRef]
  77. Guyot, D.; Giraud, F.; Simon, F.; Corgier, D.; Marvillet, C.; Tremeac, B. Building energy model calibration: A detailed case study using sub-hourly measured data. Energy Build. 2020, 223, 110189. [Google Scholar] [CrossRef]
  78. Koulamas, C.; Kalogeras, A.; Pacheco-Torres, R.; Casillas, J.; Ferrarini, L. Suitability analysis of modeling and assessment approaches in energy efficiency in buildings. Energy Build. 2018, 158, 1662–1682. [Google Scholar] [CrossRef]
  79. Manfren, M.; Aste, N.; Moshksar, R. Calibration and uncertainty analysis for computer models—A meta-model based approach for integrated building energy simulation. Appl. Energy 2013, 103, 627–641. [Google Scholar] [CrossRef]
  80. Booth, A.; Choudhary, R.; Spiegelhalter, D. A hierarchical Bayesian framework for calibrating micro-level models with macro-level data. J. Build. Perform. Simul. 2013, 6, 293–318. [Google Scholar] [CrossRef]
  81. Zhao, F.; Lee, S.H.; Augenbroe, G. Reconstructing building stock to replicate energy consumption data. Energy Build. 2016, 117, 301–312. [Google Scholar] [CrossRef]
  82. Lim, H.; Zhai, Z.J. Review on stochastic modeling methods for building stock energy prediction. Build. Simul. 2017, 10, 607–624. [Google Scholar] [CrossRef] [Green Version]
  83. Hong, S.M.; Paterson, G.; Burman, E.; Steadman, P.; Mumovic, D. A comparative study of benchmarking approaches for non-domestic buildings: Part 1—Top-down approach. Int. J. Sustain. Built Environ. 2013, 2, 119–130. [Google Scholar] [CrossRef] [Green Version]
  84. Burman, E.; Hong, S.-M.; Paterson, G.; Kimpian, J.; Mumovic, D. A comparative study of benchmarking approaches for non-domestic buildings: Part 2—Bottom-up approach. Int. J. Sustain. Built Environ. 2014, 3, 247–261. [Google Scholar] [CrossRef] [Green Version]
  85. Kneifel, J.; Webb, D. Predicting energy performance of a net-zero energy building: A statistical approach. Appl. Energy 2016, 178, 468–483. [Google Scholar] [CrossRef] [PubMed]
  86. Wang, Y.; Kuckelkorn, J.; Zhao, F.-Y.; Spliethoff, H.; Lang, W. A state of art of review on interactions between energy performance and indoor environment quality in Passive House buildings. Renew. Sustain. Energy Rev. 2017, 72, 1303–1319. [Google Scholar] [CrossRef]
  87. Deng, S.; Wang, R.; Dai, Y. How to evaluate performance of net zero energy building—A literature research. Energy 2014, 71, 1–16. [Google Scholar] [CrossRef]
  88. Berardi, U. ZEB and nZEB (definitions, design methodologies, good practices, and case studies). In Handbook of Energy Efficiency in Buildings: A Life Cycle Approach; Butterworth-Heinemann: Oxford, UK, 2018. [Google Scholar]
  89. Magrini, A.; Lentini, G.; Cuman, S.; Bodrato, A.; Marenco, L. From nearly zero energy buildings (NZEB) to positive energy buildings (PEB): The next challenge—The most recent European trends with some notes on the energy analysis of a forerunner PEB example. Dev. Built Environ. 2020, 3, 100019. [Google Scholar] [CrossRef]
  90. Zangheri, P.; Armani, R.; Pietrobon, M.; Pagliano, L. Identification of cost-optimal and NZEB refurbishment levels for representative climates and building typologies across Europe. Energy Effic. 2018, 11, 337–369. [Google Scholar] [CrossRef] [Green Version]
  91. Goel, S.; Baker, C.; Wolf, D.; Henderson, P.; Wang, N.; Rosenberg, M. A Simplified Energy Modeling Approach for Buildings (C009). In Proceedings of the 2018 Building Performance Analysis Conference and SimBuild, Chicago, IL, USA, 26–28 September 2018. [Google Scholar]
  92. Meng, Q.; Mourshed, M. Degree-day based non-domestic building energy analytics and modelling should use building and type specific base temperatures. Energy Build. 2017, 155, 260–268. [Google Scholar] [CrossRef]
  93. Kotireddy, R.; Hoes, P.-J.; Hensen, J.L. A methodology for performance robustness assessment of low-energy buildings using scenario analysis. Appl. Energy 2018, 212, 428–442. [Google Scholar] [CrossRef]
  94. Schlueter, A.; Geyer, P. Linking BIM and Design of Experiments to balance architectural and technical design factors for energy performance. Autom. Constr. 2018, 86, 33–43. [Google Scholar] [CrossRef]
  95. Jaffal, I.; Inard, C.; Ghiaus, C. Fast method to predict building heating demand based on the design of experiments. Energy Build. 2009, 41, 669–677. [Google Scholar] [CrossRef]
  96. Pernigotto, G.; Prada, A.; Gasparella, A.; Hensen, J.L.M. Development of sets of simplified building models for building simulation. In Proceedings of the 3rd International High Performance Buildings Conference, Purdue, IL, USA, 14–17 July 2014. [Google Scholar]
  97. Dogan, T.; Reinhart, C.; Michalatos, P. Autozoner: An algorithm for automatic thermal zoning of buildings with unknown interior space definitions. J. Build. Perform. Simul. 2016, 9, 176–189. [Google Scholar] [CrossRef]
  98. Dogan, T.; Reinhart, C. Shoeboxer: An algorithm for abstracted rapid multi-zone urban building energy model generation and simulation. Energy Build. 2017, 140, 140–153. [Google Scholar] [CrossRef]
  99. Delmastro, C.; Mutani, G.; Corgnati, S.P. A supporting method for selecting cost-optimal energy retrofit policies for residential buildings at the urban scale. Energy Policy 2016, 99, 42–56. [Google Scholar] [CrossRef]
  100. Ghiassi, N.; Mahdavi, A. Reductive bottom-up urban energy computing supported by multivariate cluster analysis. Energy Build. 2017, 144, 372–386. [Google Scholar] [CrossRef]
  101. Delmastro, C.; Gargiulo, M. Capturing the long-term interdependencies between building thermal energy supply and demand in urban planning strategies. Appl. Energy 2020, 268, 114774. [Google Scholar] [CrossRef]
  102. Ballarini, I.; Corrado, V. A New Methodology for Assessing the Energy Consumption of Building Stocks. Energies 2017, 10, 1102. [Google Scholar] [CrossRef] [Green Version]
  103. Goel, S.; Wang, N.; Gonzalez, J.; Horsey, H.; Long, N. Streamlining Building Efficiency Evaluation with DOE’s Asset Score Preview. In Proceedings of the 2016 ACEEE Summer Study on Energy Efficiency in Buildings, Pacific Grove, CA, USA, 21–26 August 2016. [Google Scholar]
  104. Badiei, A.; Allinson, D.; Lomas, K. Automated dynamic thermal simulation of houses and housing stocks using readily available reduced data. Energy Build. 2019, 203, 109431. [Google Scholar] [CrossRef]
  105. IEA. Annex 67—Energy Flexible Buildings; IEA: Paris, France, 2019. [Google Scholar]
  106. IEA. Annex 79—Occupant-Centric Building Design and Operation; IEA: Paris, France, 2018. [Google Scholar]
  107. Jack, R.; Loveday, D.; Allinson, D.; Lomas, K. First evidence for the reliability of building co-heating tests. Build. Res. Inf. 2018, 46, 383–401. [Google Scholar] [CrossRef] [Green Version]
  108. Lomas, K.; Oliveira, S.; Warren, P.; Haines, V.; Chatterton, T.; Beizaee, A.; Prestwood, E.; Gething, B. Do domestic heating controls save energy? A review of the evidence. Renew. Sustain. Energy Rev. 2018, 93, 52–75. [Google Scholar] [CrossRef]
  109. Mathieu, J.L.; Price, P.N.; Kiliccote, S.; Piette, M.A. Quantifying Changes in Building Electricity Use, With Application to Demand Response. IEEE Trans. Smart Grid 2011, 2, 507–518. [Google Scholar] [CrossRef] [Green Version]
  110. Jayaweera, T.; Haeri, H.; Gowans, D. The Uniform Methods Project: Methods for Determining Energy Efficiency Savings for Specific Measures. Contract 2013, 303, 275–3000. [Google Scholar]
  111. EVO IPMVP New Construction Subcommittee. International Performance Measurement and Verification Protocol: Concepts and Option for Determining Energy Savings in New Construction; Efficiency Valuation Organization (EVO): Washington, DC, USA, 2003; Volume 3. [Google Scholar]
  112. FEMP. Federal Energy Management Program, M&V Guidelines: Measurement and Verification for Federal Energy Projects Version 3.0; U.S. Department of Energy Federal Energy Management Program: Washington, DC, USA, 2008. [Google Scholar]
  113. CalTRACK Methods. Available online: http://docs.caltrack.org/en/latest/methods.html (accessed on 20 May 2020).
  114. Investor Confidence Project. Available online: https://europe.eeperformance.org/ (accessed on 31 August 2020).
  115. ISO. ISO 16346:2013 Energy Performance of Buildings—Assessment of Overall Energy Performance; ISO: Geneva, Switzerland, 2013. [Google Scholar]
  116. American Society of Heating, Refrigerating and Air-Conditioning Engineers (ASHRAE). ASHRAE Guideline 14–2014: Measurement of Energy, Demand, and Water Savings; American Society of Heating, Refrigerating and Air-Conditioning Engineers: Atlanta, GA, USA, 2014. [Google Scholar]
  117. ISO. ISO 50006:2014 Energy Management Systems—Measuring Energy Performance Using Energy Baselines (EnB) and Energy Performance Indicators (EnPI)—General Principles and Guidance; ISO: Geneva, Switzerland, 2014. [Google Scholar]
  118. Fabrizio, E.; Monetti, V. Methodologies and Advancements in the Calibration of Building Energy Models. Energies 2015, 8, 2548–2574. [Google Scholar] [CrossRef] [Green Version]
  119. Paulus, M.T. Algorithm for explicit solution to the three parameter linear change-point regression model. Sci. Technol. Built Environ. 2017, 23, 1026–1035. [Google Scholar] [CrossRef]
  120. RMV2.0-LBNL M&V2.0 Tool. Available online: https://lbnl-eta.github.io/RMV2.0/ (accessed on 20 May 2020).
  121. Masuda, H.; Claridge, D.E. Statistical modeling of the building energy balance variable for screening of metered energy use in large commercial buildings. Energy Build. 2014, 77, 292–303. [Google Scholar] [CrossRef]
  122. Lin, G.; Claridge, D.E. A temperature-based approach to detect abnormal building energy consumption. Energy Build. 2015, 93, 110–118. [Google Scholar] [CrossRef]
  123. Kohler, M.; Blond, N.; Clappier, A. A city scale degree-day method to assess building space heating energy demands in Strasbourg Eurometropolis (France). Appl. Energy 2016, 184, 40–54. [Google Scholar] [CrossRef]
  124. Fazeli, R.; Ruth, M.; Davidsdottir, B. Temperature response functions for residential energy demand—A review of models. Urban Clim. 2016, 15, 45–59. [Google Scholar] [CrossRef]
  125. Østergård, T.; Jensen, R.L.; Maagaard, S.E. A comparison of six metamodeling techniques applied to building performance simulations. Appl. Energy 2018, 211, 89–103. [Google Scholar] [CrossRef]
  126. Westermann, P.; Evins, R. Surrogate modelling for sustainable building design—A review. Energy Build. 2019, 198, 170–186. [Google Scholar] [CrossRef]
  127. Paulus, M.T.; Claridge, D.E.; Culp, C. Algorithm for automating the selection of a temperature dependent change point model. Energy Build. 2015, 87, 95–104. [Google Scholar] [CrossRef]
  128. Tronchin, L.; Manfren, M.; James, P.A. Linking design and operation performance analysis through model calibration: Parametric assessment on a Passive House building. Energy 2018, 165, 26–40. [Google Scholar] [CrossRef] [Green Version]
  129. Allard, I.; Olofsson, T.; Nair, G. Energy evaluation of residential buildings: Performance gap analysis incorporating uncertainties in the evaluation methods. Build. Simul. 2018, 11, 725–737. [Google Scholar] [CrossRef]
  130. Manfren, M.; Nastasi, B.; Tronchin, L. Linking Design and Operation Phase Energy Performance Analysis Through Regression-Based Approaches. Front. Energy Res. 2020, 8, 288. [Google Scholar] [CrossRef]
  131. Manfren, M.; Nastasi, B. Parametric Performance Analysis and Energy Model Calibration Workflow Integration—A Scalable Approach for Buildings. Energies 2020, 13, 621. [Google Scholar] [CrossRef] [Green Version]
  132. ISO. ISO 50001:2018, Energy Management Systems—Requirements with Guidance for Use; ISO: Geneva, Switzerland, 2018. [Google Scholar]
  133. Miller, C.; Nagy, Z.; Schlueter, A. Automated daily pattern filtering of measured building performance data. Autom. Constr. 2015, 49, 1–17. [Google Scholar] [CrossRef]
  134. Jalori, S.; Reddy, T.A. A new clustering method to identify outliers and diurnal schedules from building energy interval data. Ashrae Trans. 2015, 121, 33. [Google Scholar]
  135. Richard, M.A.; Fortin, H.; Poulin, A.; Leduc, M.-A.; Fournier, M. Daily load profiles clustering: A powerful tool for demand side management in medium-sized industries. In Proceedings of the ACEEE Summer Study on Energy Efficiency in Industry, Denver, CO, USA, 15–18 August 2017. [Google Scholar]
  136. Lund, P.D.; Lindgren, J.; Mikkola, J.; Salpakari, J. Review of energy system flexibility measures to enable high levels of variable renewable electricity. Renew. Sustain. Energy Rev. 2015, 45, 785–807. [Google Scholar] [CrossRef] [Green Version]
  137. Poncelet, K.; Delarue, E.; Six, D.; Duerinck, J.; D’Haeseleer, W. Impact of the level of temporal and operational detail in energy-system planning models. Appl. Energy 2016, 162, 631–643. [Google Scholar] [CrossRef] [Green Version]
  138. Adhikari, R.S.; Aste, N.; Manfren, M. Multi-commodity network flow models for dynamic energy management–Smart Grid applications. Energy Procedia 2012, 14, 1374–1379. [Google Scholar] [CrossRef] [Green Version]
  139. Manfren, M. Multi-commodity network flow models for dynamic energy management–Mathematical formulation. Energy Procedia 2012, 14, 1380–1385. [Google Scholar] [CrossRef] [Green Version]
  140. Kraning, M.; Chu, E.; Lavaei, J.; Boyd, S. Dynamic Network Energy Management via Proximal Message Passing. Found. Trends® Optim. 2014, 1, 73–126. [Google Scholar] [CrossRef] [Green Version]
  141. Dorfner, J. Open Source Modelling and Optimisation of Energy Infrastructure at Urban Scale; Technische Universität München: München, Germany, 2016. [Google Scholar]
  142. Mazzoni, S.; Ooi, S.; Nastasi, B.; Romagnoli, A. Energy storage technologies as techno-economic parameters for master-planning and optimal dispatch in smart multi energy systems. Appl. Energy 2019, 254, 113682. [Google Scholar] [CrossRef]
  143. Jentsch, M.F.; Bahaj, A.S.; James, P.A. Climate change future proofing of buildings—Generation and assessment of building simulation weather files. Energy Build. 2008, 40, 2148–2168. [Google Scholar] [CrossRef]
  144. Jentsch, M.F.; James, P.A.B.; Bourikas, L.; Bahaj, A.S. Transforming existing weather data for worldwide locations to enable energy and building performance simulation under future climates. Renew. Energy 2013, 55, 514–524. [Google Scholar] [CrossRef]
  145. Dias, J.B.; Da Graça, G.C.; Soares, P.M. Comparison of methodologies for generation of future weather data for building thermal energy simulation. Energy Build. 2020, 206, 109556. [Google Scholar] [CrossRef]
  146. Stadler, P.; Girardin, L.; Ashouri, A.; Maréchal, F. Contribution of Model Predictive Control in the Integration of Renewable Energy Sources within the Built Environment. Front. Energy Res. 2018, 6, 22. [Google Scholar] [CrossRef] [Green Version]
  147. Orehounig, K.; Mavromatidis, G.; Evins, R.; Dorer, V.; Carmeliet, J. Towards an energy sustainable community: An energy system analysis for a village in Switzerland. Energy Build. 2014, 84, 277–286. [Google Scholar] [CrossRef]
  148. Orehounig, K.; Evins, R.; Dorer, V. Integration of decentralized energy systems in neighbourhoods using the energy hub approach. Appl. Energy 2015, 154, 277–289. [Google Scholar] [CrossRef]
  149. Lammers, N.; Kissock, K.; Abels, B.; Sever, F. Measuring Progress with Normalized Energy Intensity. SAE Int. J. Mater. Manuf. 2011, 4, 460–467. [Google Scholar] [CrossRef] [Green Version]
  150. Hallinan, K.P.; Brodrick, P.; Northridge, J.; Kissock, J.K.; Brecha, R.J. Establishing Building Recommissioning Priorities and Potential Energy Savings from Utility Energy Data; ASHRAE: Peachtree Corners, GA, USA, 2011. [Google Scholar]
  151. Hallinan, K.P.; Kissock, J.K.; Brecha, R.J.; Mitchell, A. Targeting Residential Energy Reduction for City Utilities Using Historical Electrical Utility Data and Readily Available Building Data; ASHRAE: Peachtree Corners, GA, USA, 2011. [Google Scholar]
  152. Danov, S.; Carbonell, J.; Cipriano, J.; Martí-Herrero, J. Approaches to evaluate building energy performance from daily consumption data considering dynamic and solar gain effects. Energy Build. 2013, 57, 110–118. [Google Scholar] [CrossRef]
  153. Masuda, H.; Claridge, D.E. Inclusion of building envelope thermal lag effects in linear regression models of daily basis building energy use data. In Proceedings of the 12th International Conference for Enhanced Building Operations, Manchester, UK, USA, 22–26 October 2012; Available online: https://oaktrust.library.tamu.edu/handle/1969.1/148946 (accessed on 25 November 2020).
  154. Bynum, J.D.; Claridge, D.E.; Curtin, J.M. Development and testing of an Automated Building Commissioning Analysis Tool (ABCAT). Energy Build. 2012, 55, 607–617. [Google Scholar] [CrossRef]
  155. Hitchin, R.; Knight, I. Daily energy consumption signatures and control charts for air-conditioned buildings. Energy Build. 2016, 112, 101–109. [Google Scholar] [CrossRef] [Green Version]
  156. Jalori, S.; Reddy, T.A. A unified inverse modeling framework for whole-building energy interval data: Daily and hourly baseline modeling and short-term load forecasting. Ashrae Trans. 2015, 121, 156. [Google Scholar]
  157. Abushakra, B.; Paulus, M.T. An hourly hybrid multi-variate change-point inverse model using short-term monitored data for annual prediction of building energy performance, part III: Results and analysis (1404-RP). Sci. Technol. Built Environ. 2016, 22, 996–1009. [Google Scholar] [CrossRef]
  158. Bauwens, G.; Roels, S. Co-heating test: A state-of-the-art. Energy Build. 2014, 82, 163–172. [Google Scholar] [CrossRef]
  159. Erkoreka, A.; Garcia, E.; Martin, K.; Teres-Zubiaga, J.; Del Portillo, L. In-use office building energy characterization through basic monitoring and modelling. Energy Build. 2016, 119, 256–266. [Google Scholar] [CrossRef]
  160. Giraldo-Soto, C.; Erkoreka, A.; Mora, L.; Uriarte, I.; Del Portillo, L.A. Monitoring System Analysis for Evaluating a Building’s Envelope Energy Performance through Estimation of Its Heat Loss Coefficient. Sensors 2018, 18, 2360. [Google Scholar] [CrossRef] [Green Version]
  161. Uriarte, I.; Erkoreka, A.; Giraldo-Soto, C.; Martin-Escudero, K.; Uriarte, A.; Eguia, P. Mathematical development of an average method for estimating the reduction of the Heat Loss Coefficient of an energetically retrofitted occupied office building. Energy Build. 2019, 192, 101–122. [Google Scholar] [CrossRef]
  162. Busato, F.; Lazzarin, R.M.; Noro, M. Energy and economic analysis of different heat pump systems for space heating. Int. J. Low Carbon Technol. 2012, 7, 104–112. [Google Scholar] [CrossRef] [Green Version]
  163. Busato, F.; Lazzarin, R.; Noro, M. Two years of recorded data for a multisource heat pump system: A performance analysis. Appl. Therm. Eng. 2013, 57, 39–47. [Google Scholar] [CrossRef]
  164. Krese, G.; Lampret, Ž.; Butala, V.; Prek, M. Determination of a Building’s balance point temperature as an energy characteristic. Energy 2018, 165, 1034–1049. [Google Scholar] [CrossRef]
  165. Sjögren, J.U.; Andersson, S.; Olofsson, T. Sensitivity of the total heat loss coefficient determined by the energy signature approach to different time periods and gained energy. Energy Build. 2009, 41, 801–808. [Google Scholar] [CrossRef]
  166. Vesterberg, J.; Andersson, S.; Olofsson, T. Robustness of a regression approach, aimed for calibration of whole building energy simulation tools. Energy Build. 2014, 81, 430–434. [Google Scholar] [CrossRef]
  167. Meng, Q.; Xiong, C.; Mourshed, M.; Wu, M.; Ren, X.; Wang, W.; Li, Y.; Song, H. Change-point multivariable quantile regression to explore effect of weather variables on building energy consumption and estimate base temperature range. Sustain. Cities Soc. 2020, 53, 101900. [Google Scholar] [CrossRef]
  168. Oh, S.; Haberl, J.S.; Baltazar, J.-C. Analysis methods for characterizing energy saving opportunities from home automation devices using smart meter data. Energy Build. 2020, 216, 109955. [Google Scholar] [CrossRef]
  169. Westermann, P.; Deb, C.; Schlueter, A.; Evins, R. Unsupervised learning of energy signatures to identify the heating system and building type using smart meter data. Appl. Energy 2020, 264, 114715. [Google Scholar] [CrossRef]
  170. Pasichnyi, O.; Wallin, J.; Kordas, O. Data-driven building archetypes for urban building energy modelling. Energy 2019, 181, 360–377. [Google Scholar] [CrossRef]
  171. Qomi, M.J.A.; Noshadravan, A.; Sobstyl, J.M.; Toole, J.; Ferreira, J.; Pellenq, R.J.-M.; Ulm, F.-J.; Gonzalez, M.C. Data analytics for simplifying thermal efficiency planning in cities. J. R. Soc. Interface 2016, 13, 20150971. [Google Scholar] [CrossRef] [Green Version]
  172. Afshari, A.; Friedrich, L.A. Inverse modeling of the urban energy system using hourly electricity demand and weather measurements, Part 1: Black-box model. Energy Build. 2017, 157, 126–138. [Google Scholar] [CrossRef]
  173. Afshari, A.; Liu, N. Inverse modeling of the urban energy system using hourly electricity demand and weather measurements, Part 2: Gray-box model. Energy Build. 2017, 157, 139–156. [Google Scholar] [CrossRef]
  174. Catalina, T.; Virgone, J.; Blanco, E. Development and validation of regression models to predict monthly heating demand for residential buildings. Energy Build. 2008, 40, 1825–1832. [Google Scholar] [CrossRef]
  175. Hygh, J.S.; Decarolis, J.F.; Hill, D.B.; Ranjithan, S.R. Multivariate regression as an energy assessment tool in early building design. Build. Environ. 2012, 57, 165–175. [Google Scholar] [CrossRef]
  176. Asadi, S.; Amiri, S.S.; Mottahedi, M. On the development of multi-linear regression analysis to assess energy consumption in the early stages of building design. Energy Build. 2014, 85, 246–255. [Google Scholar] [CrossRef]
  177. Al Gharably, M.; Decarolis, J.F.; Ranjithan, S.R. An enhanced linear regression-based building energy model (LRBEM+) for early design. J. Build. Perform. Simul. 2015, 9, 115–133. [Google Scholar] [CrossRef]
  178. Ipbüker, C.; Valge, M.; Kalbe, K.; Mauring, T.; Tkaczyk, A. Case Study of Multiple Regression as Evaluation Tool for the Study of Relationships between Energy Demand, Air Tightness, and Associated Factors. J. Energy Eng. 2017, 143, 04016027. [Google Scholar] [CrossRef]
  179. Ligier, S.; Robillart, M.; Schalbart, P.; Peuportier, B. Energy Performance Contracting Methodology Based upon Simulation and Measurement. In Proceedings of the Building Simulation 2017, San Francisco, CA, USA, 7–9 August 2017. [Google Scholar]
  180. Abdelalim, A.; O’Brien, W.; Shi, Z. Data visualization and analysis of energy flow on a multi-zone building scale. Autom. Constr. 2017, 84, 258–273. [Google Scholar] [CrossRef]
  181. Liu, X.; Mancarella, P. Modelling, assessment and Sankey diagrams of integrated electricity-heat-gas networks in multi-vector district energy systems. Appl. Energy 2016, 167, 336–352. [Google Scholar] [CrossRef]
  182. Masuda, H.; Claridge, D. Estimation of Building Parameters Using Simplified Energy Balance Model and Metered Whole Building Energy Use; Texas A&M University Libraries: Killeen, TX, USA, 2012. [Google Scholar]
  183. Tronchin, L.; Manfren, M.; Tagliabue, L.C. Optimization of building energy performance by means of multi-scale analysis—Lessons learned from case studies. Sustain. Cities Soc. 2016, 27, 296–306. [Google Scholar] [CrossRef]
  184. Tronchin, L.; Manfren, M.; Nastasi, B. Energy analytics for supporting built environment decarbonisation. Energy Procedia 2019, 157, 1486–1493. [Google Scholar] [CrossRef]
  185. Manfren, M.; Nastasi, B. From in-situ measurement to regression and time series models: An overview of trends and prospects for building performance modelling. AIP Conf. Proc. 2019, 2123, 20100. [Google Scholar] [CrossRef]
  186. ISO. ISO 52000-1:2017, Energy Performance of Buildings—Overarching EPB Assessment–Part 1: General Framework and Procedures; ISO: Geneva, Switzerland, 2017. [Google Scholar]
  187. Li, F.G.N.; Smith, A.; Biddulph, P.; Hamilton, I.; Lowe, R.J.; Mavrogianni, A.; Oikonomou, E.; Raslan, R.; Stamp, S.; Stone, A.; et al. Solid-wallU-values: Heat flux measurements compared with standard assumptions. Build. Res. Inf. 2015, 43, 238–252. [Google Scholar] [CrossRef] [Green Version]
  188. Rasooli, A.; Itard, L. In-situ characterization of walls’ thermal resistance: An extension to the ISO 9869 standard method. Energy Build. 2018, 179, 374–383. [Google Scholar] [CrossRef]
  189. Alzetto, F.; Meulemans, J.; Pandraud, G.; Roux, D. A perturbation method to estimate building thermal performance. Comptes Rendus Chim. 2018, 21, 938–942. [Google Scholar] [CrossRef]
  190. Meulemans, J. An Assessment of the QUB/e Method for Fast In Situ Measurements of the Thermal Performance of Building Fabrics in Cold Climates. In Cold Climate HVAC Conference; Springer: Berlin/Heidelberg, Germany, 2018; pp. 317–326. [Google Scholar]
  191. Ahmad, N.; Ghiaus, C.; Thiery, T. Influence of Initial and Boundary Conditions on the Accuracy of the QUB Method to Determine the Overall Heat Loss Coefficient of a Building. Energies 2020, 13, 284. [Google Scholar] [CrossRef] [Green Version]
  192. Bouchié, R.; Alzetto, F.; Brun, A.; Boisson, P.; Thebault, S. Short methodologies for in-situ assessment ofthe intrinsic thermal performance of the building envelope. In Proceedings of the Sustainable Places 2014 (SP2014), Nice, France, 1–3 October 2014. [Google Scholar]
  193. Thébault, S.; Bouchié, R. Refinement of the ISABELE method regarding uncertainty quantification and thermal dynamics modelling. Energy Build. 2018, 178, 182–205. [Google Scholar] [CrossRef]
  194. Michalak, P. A thermal network model for the dynamic simulation of the energy performance of buildings with the time varying ventilation flow. Energy Build. 2019, 202, 109337. [Google Scholar] [CrossRef]
  195. Lundström, L.; Akander, J.; Zambrano, J. Development of a Space Heating Model Suitable for the Automated Model Generation of Existing Multifamily Buildings—A Case Study in Nordic Climate. Energies 2019, 12, 485. [Google Scholar] [CrossRef] [Green Version]
  196. Fuentenueva, C.; DeNaveros, I.; Ghiaus, C.; Ordoñez, J.; Ruiz, D.P. Thermal networks considering graph theory and thermodynamics. In Proceedings of the 12th International Conference on Heat Transfer, Fluid Mechanics and Thermodynamics, Malaga, Spain, 11–13 July 2016; pp. 1568–1573. [Google Scholar]
  197. Naveros, I.; Ghiaus, C.; Ordoñez, J.; Ruíz, D.P. Thermal networks from the heat equation by using the finite element method. WIT Trans. Eng. Sci. 2016, 106, 33–43. [Google Scholar] [CrossRef] [Green Version]
  198. Ghiaus, C.; Ahmad, N. Thermal circuits assembling and state-space extraction for modelling heat transfer in buildings. Energy 2020, 195, 117019. [Google Scholar] [CrossRef]
  199. Ramallo-González, A.P.; Eames, M.E.; Natarajan, S.; Fosas-De-Pando, D.; Coley, D.A. An analytical heat wave definition based on the impact on buildings and occupants. Energy Build. 2020, 216, 109923. [Google Scholar] [CrossRef]
  200. Michalak, P. The development and validation of the linear time varying Simulink-based model for the dynamic simulation of the thermal performance of buildings. Energy Build. 2017, 141, 333–340. [Google Scholar] [CrossRef]
  201. Hong, T.; Lee, S.H. Integrating physics-based models with sensor data: An inverse modeling approach. Build. Environ. 2019, 154, 23–31. [Google Scholar] [CrossRef] [Green Version]
  202. Juricic, S.; Goffart, J.; Rouchier, S.; Foucquier, A.; Cellier, N.; Fraisse, G. Influence of weather natural variability on the thermal characterisation of a building envelope. arXiv 2020, arXiv:2011.13593. [Google Scholar]
  203. Baasch, G.; Wicikowski, A.; Faure, G.; Evins, R. Comparing Gray Box Methods to Derive Building Properties from Smart Thermostat Data. In Proceedings of the 6th ACM International Conference on Systems for Energy-Efficient Buildings, Cities, and Transportation, New York, NY, USA, 13–14 November 2019; ACM: New York, NY, USA, 2019; pp. 223–232. [Google Scholar]
  204. Raillon, L.; Ghiaus, C. Study of Error Propagation in the Transformations of Dynamic Thermal Models of Buildings. J. Control. Sci. Eng. 2017, 2017, 1–15. [Google Scholar] [CrossRef] [Green Version]
  205. Gustin, M.; McLeod, R.S.; Lomas, K.J. Forecasting indoor temperatures during heatwaves using time series models. Build. Environ. 2018, 143, 727–739. [Google Scholar] [CrossRef]
  206. Gustin, M.; McLeod, R.S.; Lomas, K.J. Can semi-parametric additive models outperform linear models, when forecasting indoor temperatures in free-running buildings? Energy Build. 2019, 193, 250–266. [Google Scholar] [CrossRef]
  207. Panão, M.J.O.; Santos, C.A.; Mateus, N.M.; Da Graça, G.C. Validation of a lumped RC model for thermal simulation of a double skin natural and mechanical ventilated test cell. Energy Build. 2016, 121, 92–103. [Google Scholar] [CrossRef]
  208. Naveros-Mesa, I.; Ghiaus, C.; Ruiz, D.; Castaño, S. Physical parameters identification of walls using ARX models obtained by deduction. Energy Build. 2015, 108, 317–329. [Google Scholar] [CrossRef]
  209. Andriamamonjy, A.; Klein, R.; Saelens, D. Automated grey box model implementation using BIM and Modelica. Energy Build. 2019, 209–225. [Google Scholar] [CrossRef]
  210. Lehmann, B.; Gyalistras, D.; Gwerder, M.; Wirth, K.; Carl, S. Intermediate complexity model for Model Predictive Control of Integrated Room Automation. Energy Build. 2013, 58, 250–262. [Google Scholar] [CrossRef]
  211. Raillon, L.; Ghiaus, C. An efficient Bayesian experimental calibration of dynamic thermal models. Energy 2018, 152, 818–833. [Google Scholar] [CrossRef] [Green Version]
  212. Raillon, L.L.; Ghiaus, C. Sequential Monte Carlo for states and parameters estimation in dynamic thermal models. In Proceedings of the Building Simulation 2017 Conference, San Francisco, CA, USA, 7–9 August 2017; pp. 988–997. [Google Scholar]
  213. Serale, G.; Fiorentini, M.; Capozzoli, A.; Bernardini, D.; Bemporad, A. Model Predictive Control (MPC) for Enhancing Building and HVAC System Energy Efficiency: Problem Formulation, Applications and Opportunities. Energies 2018, 11, 631. [Google Scholar] [CrossRef] [Green Version]
  214. Ahmad, M.W.; Mourshed, M.; Mundow, D.; Sisinni, M.; Rezgui, Y. Building energy metering and environmental monitoring—A state-of-the-art review and directions for future research. Energy Build. 2016, 120, 85–102. [Google Scholar] [CrossRef] [Green Version]
  215. Carstens, H.; Xia, X.; Yadavalli, S. Measurement uncertainty in energy monitoring: Present state of the art. Renew. Sustain. Energy Rev. 2018, 82, 2791–2805. [Google Scholar] [CrossRef]
  216. Senave, M.; Roels, S.; Verbeke, S.; Saelens, D. Analysis of the influence of the definition of the interior dwelling temperature on the characterization of the heat loss coefficient via on-board monitoring. Energy Build. 2020, 215, 109860. [Google Scholar] [CrossRef]
  217. Lydon, G.; Caranovic, S.; Hischier, I.; Schlueter, A. Coupled simulation of thermally active building systems to support a digital twin. Energy Build. 2019, 202, 109298. [Google Scholar] [CrossRef]
  218. Hamilton, I.; Summerfield, A.; Lowe, R.J.; Ruyssevelt, P.; Elwell, C.A.; Oreszczyn, T. Energy epidemiology: A new approach to end-use energy demand research. Build. Res. Inf. 2013, 41, 482–497. [Google Scholar] [CrossRef] [Green Version]
  219. Hamilton, I.; Summerfield, A.; Oreszczyn, T.; Ruyssevelt, P. Using epidemiological methods in energy and buildings research to achieve carbon emission targets. Energy Build. 2017, 154, 188–197. [Google Scholar] [CrossRef]
  220. Junker, R.G.; Azar, A.G.; Lopes, R.; Lindberg, K.B.; Reynders, G.; Relan, R.; Madsen, H. Characterizing the energy flexibility of buildings and districts. Appl. Energy 2018, 225, 175–182. [Google Scholar] [CrossRef]
  221. Kohlhepp, P.; Harb, H.; Wolisz, H.; Waczowicz, S.; Müller, D.; Hagenmeyer, V. Large-scale grid integration of residential thermal energy storages as demand-side flexibility resource: A review of international field studies. Renew. Sustain. Energy Rev. 2019, 101, 527–547. [Google Scholar] [CrossRef]
  222. Vijay, A.; Hawkes, A. Demand side flexibility from residential heating to absorb surplus renewables in low carbon futures. Renew. Energy 2019, 138, 598–609. [Google Scholar] [CrossRef]
  223. Bee, E.; Prada, A.; Baggio, P.; Psimopoulos, E. Air-source heat pump and photovoltaic systems for residential heating and cooling: Potential of self-consumption in different European climates. Build. Simul. 2019, 12, 453–463. [Google Scholar] [CrossRef]
  224. Psimopoulos, E.; Bee, E.; Widén, J.; Bales, C. Techno-economic analysis of control algorithms for an exhaust air heat pump system for detached houses coupled to a photovoltaic system. Appl. Energy 2019, 249, 355–367. [Google Scholar] [CrossRef]
  225. Facci, A.L.; Krastev, V.K.; Falcucci, G.; Ubertini, S. Smart integration of photovoltaic production, heat pump and thermal energy storage in residential applications. Sol. Energy 2019, 192, 133–143. [Google Scholar] [CrossRef]
  226. Oliveira-Lima, J.A.; Delgado-Gomes, V.; Martins, J.; Lima, C. Standard-based service-oriented infrastructure to integrate intelligent buildings in distributed generation and smart grids. Energy Build. 2014, 76, 450–458. [Google Scholar] [CrossRef]
  227. Sun, M.; Djapic, P.; Aunedi, M.; Pudjianto, D.; Strbac, G. Benefits of smart control of hybrid heat pumps: An analysis of field trial data. Appl. Energy 2019, 247, 525–536. [Google Scholar] [CrossRef]
  228. Gui, E.M.; MacGill, I. Consumer-Centric Service Innovations in an Era of Self-Selecting Customers. Consumer, Prosumer, Prosumager: How Service Innovations will Disrupt the Utility Business Model; Academic Press: Cambridge, MA, USA, 2019; pp. 127–151. [Google Scholar] [CrossRef]
  229. Clauß, J.; Finck, C.; Vogler-Finck, P.; Beagon, P. Control strategies for building energy systems to unlock demand side flexibility—A review. In Proceedings of the IBPSA Building Simulation 2017, San Francisco, CA, USA, 7–9 August 2017. [Google Scholar]
  230. Péan, T.; Salom, J.; Costa-Castelló, R. Review of control strategies for improving the energy flexibility provided by heat pump systems in buildings. J. Process. Control. 2019, 74, 35–49. [Google Scholar] [CrossRef]
  231. Clauß, J.; Georges, L. Model complexity of heat pump systems to investigate the building energy flexibility and guidelines for model implementation. Appl. Energy 2019, 255, 113847. [Google Scholar] [CrossRef]
  232. Drgoňa, J.; Arroyo, J.; Figueroa, I.C.; Blum, D.H.; Arendt, K.; Kim, D.; Ollé, E.P.; Oravec, J.; Wetter, M.; Vrabie, D.L.; et al. All you need to know about model predictive control for buildings. Annu. Rev. Control. 2020, 50, 190–232. [Google Scholar] [CrossRef]
  233. De Coninck, R.; Baetens, R.; Saelens, D.; Woyte, A.; Helsen, L. Rule-based demand-side management of domestic hot water production with heat pumps in zero energy neighbourhoods. J. Build. Perform. Simul. 2013, 7, 271–288. [Google Scholar] [CrossRef]
  234. Klein, K.; Kalz, D.; Herkel, S. Grid impact of a net zero energy building with BiPV using different energy management strategies. In Proceedings of the International Conference CISBAT 2015 Future Buildings and Districts Sustainability from Nano to Urban Scale, Lausanne, Switzerland, 9–11 September 2015; pp. 579–584. [Google Scholar]
  235. Le Dréau, J.; Heiselberg, P. Energy flexibility of residential buildings using short term heat storage in the thermal mass. Energy 2016, 111, 991–1002. [Google Scholar] [CrossRef]
  236. Dar, U.I.; Sartori, I.; Georges, L.; Novakovic, V. Advanced control of heat pumps for improved flexibility of Net-ZEB towards the grid. Energy Build. 2014, 69, 74–84. [Google Scholar] [CrossRef]
  237. Reynders, G.; Diriken, J.; Saelens, D. A generic quantification method for the active demand response potential of structural storage in buildings. In Proceedings of the 14th International Conference of IBPSA-Building Simulation, Hyderabad, India, 7–9 December 2015; pp. 1986–1993. [Google Scholar]
  238. Turner, W.J.N.; Walker, I.S.; Roux, J. Peak load reductions: Electric load shifting with mechanical pre-cooling of residential buildings with low thermal mass. Energy 2015, 82, 1057–1067. [Google Scholar] [CrossRef]
  239. Esfehani, H.H.; Kriegel, M.; Madani, H. Load balancing potential of ground source heat pump system coupled with thermal energy storage: A Case Study for Berlin. In Proceedings of the CLIMA 2016, 12th REHVA World Congress, Aalborg, Denmark, 22–25 May 2016. [Google Scholar]
  240. Alimohammadisagvand, B.; Jokisalo, J.; Kilpeläinen, S.; Ali, M.; Sirén, K. Cost-optimal thermal energy storage system for a residential building with heat pump heating and demand response control. Appl. Energy 2016, 174, 275–287. [Google Scholar] [CrossRef]
  241. Salpakari, J.; Lund, P. Optimal and rule-based control strategies for energy flexibility in buildings with PV. Appl. Energy 2016, 161, 425–436. [Google Scholar] [CrossRef] [Green Version]
  242. Masy, G.; Georges, E.; Verhelst, C.; Lemort, V.; André, P. Smart grid energy flexible buildings through the use of heat pumps and building thermal mass as energy storage in the Belgian context. Sci. Technol. Built Environ. 2015, 21, 800–811. [Google Scholar] [CrossRef]
  243. Panão, M.J.O.; Mateus, N.M.; Da Graça, G.C. Measured and modeled performance of internal mass as a thermal energy battery for energy flexible residential buildings. Appl. Energy 2019, 239, 252–267. [Google Scholar] [CrossRef]
  244. Vivian, J.; Chiodarelli, U.; Emmi, G.; Zarrella, A. A sensitivity analysis on the heating and cooling energy flexibility of residential buildings. Sustain. Cities Soc. 2020, 52, 101815. [Google Scholar] [CrossRef]
  245. De Coninck, R.; Helsen, L. Quantification of flexibility in buildings by cost curves—Methodology and application. Appl. Energy 2016, 162, 653–665. [Google Scholar] [CrossRef]
  246. Halvgaard, R.; Poulsen, N.K.; Madsen, H.; Jorgensen, J.B. Economic Model Predictive Control for building climate control in a Smart Grid. In Proceedings of the 2012 IEEE PES Innovative Smart Grid Technologies (ISGT), Washington, DC, USA, 16–20 January 2012; pp. 1–6. [Google Scholar]
  247. Maasoumy Haghighi, M. Controlling Energy-Efficient Buildings in the Context of Smart Grid: A Cyber Physical System Approach; Technical Report No. UCB/EECS-2013-244; University of California: Berkley, CA, USA, 2014. [Google Scholar]
  248. Corbin, C.D.; Henze, G. Predictive control of residential HVAC and its impact on the grid. Part I: Simulation framework and models. J. Build. Perform. Simul. 2016, 10, 294–312. [Google Scholar] [CrossRef]
  249. Corbin, C.; Henze, G. Predictive control of residential HVAC and its impact on the grid. Part II: Simulation studies of residential HVAC as a supply following resource. J. Build. Perform. Simul. 2017, 10, 365–377. [Google Scholar] [CrossRef]
  250. Lindelöf, D.; Afshari, H.; Alisafaee, M.; Biswas, J.; Caban, M.; Mocellin, X.; Viaene, J. Field tests of an adaptive, model-predictive heating controller for residential buildings. Energy Build. 2015, 99, 292–302. [Google Scholar] [CrossRef]
  251. Garnier, A.; Eynard, J.; Caussanel, M.; Grieu, S. Predictive control of multizone heating, ventilation and air-conditioning systems in non-residential buildings. Appl. Soft Comput. 2015, 37, 847–862. [Google Scholar] [CrossRef]
  252. Kandler, C.; Wimmer, P.; Honold, J. Predictive Control and Regulation Strategies of Air-to-Water Heat Pumps. Energy Procedia 2015, 78, 2088–2093. [Google Scholar] [CrossRef] [Green Version]
  253. Blum, D.H.; Arendt, K.; Rivalin, L.; Piette, M.; Wetter, M.; Veje, C. Practical factors of envelope model setup and their effects on the performance of model predictive control for building heating, ventilating, and air conditioning systems. Appl. Energy 2019, 236, 410–425. [Google Scholar] [CrossRef]
  254. Dong, B.; Yan, D.; Li, Z.; Jin, Y.; Feng, X.; Fontenot, H. Modeling occupancy and behavior for better building design and operation—A critical review. Build. Simul. 2018, 11, 899–921. [Google Scholar] [CrossRef]
  255. Naylor, S.; Gillott, M.; Lau, T. A review of occupant-centric building control strategies to reduce building energy use. Renew. Sustain. Energy Rev. 2018, 96, 1–10. [Google Scholar] [CrossRef]
  256. Carpino, C.; Mora, D.; Arcuri, N.; De Simone, M. Behavioral variables and occupancy patterns in the design and modeling of Nearly Zero Energy Buildings. Build. Simul. 2017, 22, 860–888. [Google Scholar] [CrossRef]
  257. Caucheteux, A.; Sabar, A.E.; Boucher, V. Occupancy measurement in building: A litterature review, application on an energy efficiency research demonstrated building. Int. J. Metrol. Qual. Eng. 2013, 4, 135–144. [Google Scholar] [CrossRef] [Green Version]
  258. Naspi, F.; Arnesano, M.; Stazi, F.; D’Orazio, M.; Revel, G.M. Measuring Occupants’ Behaviour for Buildings’ Dynamic Cosimulation. J. Sens. 2018, 2018, 2756542. [Google Scholar] [CrossRef] [Green Version]
  259. Cecconi, F.R.; Manfren, M.; Tagliabue, L.C.; Ciribini, A.L.C.; De Angelis, E. Probabilistic behavioral modeling in building performance simulation: A Monte Carlo approach. Energy Build. 2017, 148, 128–141. [Google Scholar] [CrossRef]
  260. Tagliabue, L.C.; Manfren, M.; Ciribini, A.L.C.; De Angelis, E. Probabilistic behavioural modeling in building performance simulation—The Brescia eLUX lab. Energy Build. 2016, 128, 119–131. [Google Scholar] [CrossRef]
  261. De Menezes, A.C.K.; Cripps, A.; Bouchlaghem, D.; Buswell, R.A. Predicted vs. actual energy performance of non-domestic buildings: Using post-occupancy evaluation data to reduce the performance gap. Appl. Energy 2012, 97, 355–364. [Google Scholar] [CrossRef] [Green Version]
  262. Aragon, V.; Gauthier, S.; Warren, P.; James, P.A.B.; Anderson, B. Developing English domestic occupancy profiles. Build. Res. Inf. 2019, 47, 375–393. [Google Scholar] [CrossRef] [Green Version]
  263. Zangheri, P.; Pagliano, L.; Armani, R. How the comfort requirements can be used to assess and design low energy buildings: Testing the EN 15251 comfort evaluation procedure in 4 buildings. In Proceedings of the ECEEE 2011 Summer Study “Energy Efficiency First: The Foundation of a Low-Carbon Society”, Hyeres, France, 6–11 June 2011; pp. 1569–1579. [Google Scholar]
  264. Fabbri, K.; Tronchin, L. Indoor Environmental Quality in Low Energy Buildings. Energy Procedia 2015, 78, 2778–2783. [Google Scholar] [CrossRef] [Green Version]
  265. Manfren, M.; Nastasi, B.; Piana, E.A.; Tronchin, L. On the link between energy performance of building and thermal comfort: An example. AIP Conf. Proc. 2019, 2123, 20066. [Google Scholar] [CrossRef]
  266. Sarfraz, O.; Bach, C.K. Equipment power consumption and load factor profiles for buildings’ energy simulation (ASHRAE 1742-RP). Sci. Technol. Built Environ. 2018, 24, 1054–1063. [Google Scholar] [CrossRef]
  267. Gunay, B.; Shen, W.; Yang, C. Characterization of a building’s operation using automation data: A review and case study. Build. Environ. 2017, 118, 196–210. [Google Scholar] [CrossRef] [Green Version]
  268. Saini, J.; Dutta, M.; Marques, G. Indoor Air Quality Monitoring Systems Based on Internet of Things: A Systematic Review. Int. J. Environ. Res. Public Health 2020, 17, 4942. [Google Scholar] [CrossRef]
  269. Martín-Garín, A.; Millán-García, J.; Baïri, A.; Millán-Medel, J.; Sala-Lizarraga, J. Environmental monitoring system based on an Open Source Platform and the Internet of Things for a building energy retrofit. Autom. Constr. 2018, 87, 201–214. [Google Scholar] [CrossRef]
  270. Lucchi, E. Environmental Risk Management for Museums in Historic Buildings through an Innovative Approach: A Case Study of the Pinacoteca di Brera in Milan (Italy). Sustainability 2020, 12, 5155. [Google Scholar] [CrossRef]
  271. Pauwels, P.; Zhang, S.; Lee, Y.-C. Semantic web technologies in AEC industry: A literature overview. Autom. Constr. 2017, 73, 145–165. [Google Scholar] [CrossRef]
  272. Corry, E.; Pauwels, P.; Hu, S.; Keane, M.; O’Donnell, J. A performance assessment ontology for the environmental and energy management of buildings. Autom. Constr. 2015, 57, 249–259. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Diagram synthesizing the research process.
Figure 1. Diagram synthesizing the research process.
Energies 14 00679 g001
Table 1. Knowledge domain, keywords and criteria for literature selection.
Table 1. Knowledge domain, keywords and criteria for literature selection.
Domain of
Interest
Domain and KeywordsSources in WoS DatabaseSources in
Categories:
Architecture
Construction
Planning
Motivation for Criteria SelectionSource in Final Selection
“Building Energy Performance”“Building Energy Performance”
AND
“Building stock”
1335870Building energy modelling for energy planning and policy targets, utility scale analysis, parametric building performance studies.52
“Building Energy Performance”
AND
“Uncertainty”
1551705Methods based on M&V and M&T principles that can help tracking energy performance transparently (and reducing uncertainty) and that can be applied at multiple temporal and spatial scales.123
“Building Energy Performance”
AND
“Flexibility”
1027237Strategies to control buildings and enhance their energy flexibility strategies in relation to energy demand in end-uses and user behaviour.68
Table 2. Additional criteria introduced for energy modelling literature selection.
Table 2. Additional criteria introduced for energy modelling literature selection.
CriteriaDescriptionMotivation for Criteria Selection
Empirical GroundingBased on empirical data, and tested on a relevant number of cases.Reducing risk of investment in energy transitions and ensure the credibility of policies by means of evidence.
HarmonizationMethodologies in which redundancies and overlapping features are removed, ideally based on protocols and standard.Avoid redundancy, multiplication of efforts and unnecessary increase of complexity of procedures. Streamline the implementation of models and procedures.
ScalabilityCapability of analysing problems at multiple temporal and spatial scales.Ability to work coherently and consistently on multiple temporal and spatial scales.
InterpretabilityAbility to detect relevant cause-effect relationship, ideally combing statistical analysis techniques with physical understanding of phenomena.Physical interpretation can help extract insights that are fundamental for the continuous improvements of processes and technologies.
Re-configurabilityAble to be used in multiple stages of the building life-cycle, for example for design and operation, sharing similar underlying principles.Creating a certain degree of continuity in the data analysis workflow during the life-cycle of projects.
Table 3. Building energy performance analysis—Target and spatial scale of analysis.
Table 3. Building energy performance analysis—Target and spatial scale of analysis.
SourceYearTarget of AnalysisSpatial Scale of Analysis
Energy Planning and PolicyUtility Level StudyParametric Building AnalysisNationalRegionalUrbanStock
Deru et al. [68]2011
Thornton et al. [70]2011
Goel et al. [69]2011
Ballarini et al. [102]2017
Delmastro et al. [99]2016
Ghiassi et al. [100]2017
Delmastro et al. [101]2020
Goel et al. [91]2018
Meng et al. [92]2017
Pernigotto et al. [96]2014
Dogan et al. [97]2016
Dogan et al. [98]2016
Goel et al. [103]2016
Badiei et al. [104]2019
Table 4. Harmonized regression-based modelling approaches for building performance analysis.
Table 4. Harmonized regression-based modelling approaches for building performance analysis.
SourceYearEnd-UseEnergy ServicesTemporal ScaleSpatial ScaleInterpretationPhase
ResidentialNon-residentialHeatingCoolingDHWAppliancesMonthlyDailyHourlyBuilding fabricTechnical systemsWhole buildingBuilding stockCommunityPhysical approximatedDesignOperation
Lammers et al. [149]2011
Hallinan et al. [150]2011
Hallinan et al. [151]2011
Danov et al. [152]2011
Masuda and Claridge [153]2012
Bynum et al. [154]2012
Masuda and Claridge [121]2014
Paulus et al. [127]2015
Lin and Claridge [122]2015
Hitchin and Knight [155]2016
Jalori and Reddy [156]2015
Paulus [119]2017
Abushakra and Paulus [157]2016
Bauwens and Roels [158]2014
Erkoreka et al. [159]2016
Giraldo-Soto et al. [160]2018
Uriarte et al. [161]2019
Busato et al. [162]2012
Busato et al. [163]2013
Krese et al. [164]2018
Sjögren et al. [165]2009
Vesterberg et al. [166]2014
Meng and Mourshed [92]2017
Meng et al. [167]2020
Oh et al. [168]2020
Westermann et al. [169]2020
Pasichnyi et al. [170]2019
Qomi et al. [171]2016
Afshari et al. [172]2017
Afshari et al. [173]2017
Allard et al. [129]2018
Tronchin et al. [128]2018
Manfren and Nastasi [131]2020
Catalina et al. [174]2008
Hygh et al. [175]2012
Asadi et al. [176]2014
Al Gharably et al. [177]2016
Ipbüker et al. [178]2016
Goel et al. [103]2016
Table 5. Experimental protocols and applications.
Table 5. Experimental protocols and applications.
SourceYearType of Experimental ProtocolApplicationData Acquisition
ISO 9869Co-heatingQUBISABELEComponentZoneTime IntervalLength of Data Acquisition
Francis et al. [187]2015 Subhourly72 h
Rasooli and Itard [188]2018 Subhourly72 h
Erkoreka et al. [159]2016 Subhourly72 h, multiple periods
Uriarte et al. [161]2019 Subhourly72 h multiple periods
Bauwens et al. [158]2014 Daily2/3 weeks
Jack et al. [107]2017 Daily2/3 weeks
Alzetto et al. [189]2018 Subhourly1 night
Meulemans [190]2018 Subhourly1 night
Ahmad et al. [191]2019 Subhourly1 night
Rémi et al. [192]2014 Subhourly5–15 days
Thébault et al. [193]2018 Subhourly4 days
Table 6. Control strategies aimed at building flexibility for different end-uses and services.
Table 6. Control strategies aimed at building flexibility for different end-uses and services.
SourceYearEnd-usesEnergy servicesControl ObjectiveControl TypeTime ScheduleSet-pointsComfort ConstraintsLoad (Demand)Production (On-Site)Grid Connection (import/export)Tariff
ResidentialNon-residentialHeatingCoolingDHWAppliances
De Coninck et al. [233]2014 LS, ORMRBC
Klein et al. [234]2015 LS, ORMRBC
Le Dréau and Heiselberg [235]2016 LSRBC
Dar et al. [236]2014 LS, ORMRBC
Reynders et al. [237]2015 LSRBC
Turner et al. [238]2015 LSRBC
Esfehani et al. [239]2016 LS, ORMRBC
Alimohammadisagvand et al. [240]2016 LSRBC
Salpakari and Lund [241]2016 LS, ORMRBC, OC
Masy et al. [242]2015 LSRBC, OC
Psimopoulos et al. [224]2019 LSRBC
Bee et al. [223]2019 LSRBC
Oliveira Panão et al. [243]2019 LSRBC
Vivian et al. [244]2020 LSRBC
De Coninck and Helsen [245]2016 LSOC
Halvgaard et al. [246]2012 LSMPC
Maasoumy Haghighi [247]2013 LSMPC
Corbin and Henze [248]2017 LSMPC
Corbin and Henze [249]2017 LS, ORMMPC
Lindelöf et al. [250]2015 LSMPC
Garnier et al. [251]2015 LSMPC
Kandler et al. [252]2015 LS, ORMMPC
Blum et al. [253]2019 LSMPC
Table 7. Articles at the intersection of levels of analysis.
Table 7. Articles at the intersection of levels of analysis.
SourceLevelYearPattern IdentifiedPaper Title
Calleja Rodríguez et al. [75]#1#22013Reference building approach and parametric modellingUK office buildings archetypal model as methodological approach in development of regression models for predicting building energy consumption from heating and cooling demands
Goel et al. [103]#1#22016Reference building approach and parametric modellingStreamlining Building Efficiency Evaluation with DOE’s Asset Score Preview
Zhao et al. [81]#1#22016Reference building approach and parametric modellingReconstructing building stock to replicate energy consumption data
Lim et al. [82]#1#22017Reference building approach and parametric modellingReview on stochastic modeling methods for building stock energy prediction
Booth et al. [80]#1#22013Multi-level calibrationA hierarchical bayesian framework for calibrating micro-level models with macro-level data
Yang and Becerik-Gerber [74]#1#22015Multi-level calibrationA model calibration framework for simultaneous multi-level building energy simulation
Fabrizio et al. [118]#2#32015Multi-level calibrationMethodologies and advancements in the calibration of building energy models
Guyot et al. [77]#1#22020Multi-level calibrationBuilding energy model calibration: A detailed case study using sub-hourly measured data
Jalori et al. [134]#2#32015Regression-based approaches at multiple temporal and spatial scale of analysisA new clustering method to identify outliers and diurnal schedules from building energy interval data
Jalori et al. [156]#2#32015Regression-based approaches at multiple temporal and spatial scale of analysisA unified inverse modeling framework for whole-building energy interval data: Daily and hourly baseline modeling and short-term load forecasting
Ligier et al. [179]#2#32017Regression-based approaches at multiple temporal and spatial scale of analysisEnergy Performance Contracting Methodology Based upon Simulation and Measurement
Meng et al. [92]#1#22017Regression-based approaches at multiple temporal and spatial scale of analysisDegree-day based non-domestic building energy analytics and modelling should use building and type specific base temperatures
Gaetani et al. [73]#1#32016User behavioural analysis Occupant behavior in building energy simulation: Towards a fit-for-purpose modeling strategy
IEA-EBC [106]#1#32017User behavioural analysisIEA EBC-Annex 79-Occupant-Centric Building Design and Operation
IEA-EBC [105]#1#32014Flexibility and dynamic interaction with infrastructuresEBC Annex 67 Energy Flexible Buildings
Lund et al. [136]#1#32015Flexibility and dynamic interaction with infrastructuresReview of energy system flexibility measures to enable high levels of variable renewable electricity
Dominkovic et al. [25]#1#32020Flexibility and dynamic interaction with infrastructuresImplementing flexibility into energy planning models: Soft-linking of a high-level energy planning model and a short-term operational model
Ahmad et al. [214]#2#32016Automation systems, measurements, sensorsBuilding energy metering and environmental monitoring—A state-of-the-art review and directions for future research
Aste et al. [72]#1#2#32017Automation systems, measurements, sensorsBuilding Automation and Control Systems and performance optimization: A framework for analysis
Carstens et al. [215]#2#32018Automation systems, measurements, sensorsMeasurement uncertainty in energy monitoring: Present state of the art
Giraldo-Soto et al. [160]#2#32018Automation systems, measurements, sensorsMonitoring system analysis for evaluating a building’s envelope energy performance through estimation of its heat loss coefficient
Serale et al. [213]#2#32018Automation systems, measurements, sensorsModel Predictive Control (MPC) for enhancing building and HVAC system energy efficiency: Problem formulation, applications and opportunities
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Manfren, M.; Sibilla, M.; Tronchin, L. Energy Modelling and Analytics in the Built Environment—A Review of Their Role for Energy Transitions in the Construction Sector. Energies 2021, 14, 679. https://doi.org/10.3390/en14030679

AMA Style

Manfren M, Sibilla M, Tronchin L. Energy Modelling and Analytics in the Built Environment—A Review of Their Role for Energy Transitions in the Construction Sector. Energies. 2021; 14(3):679. https://doi.org/10.3390/en14030679

Chicago/Turabian Style

Manfren, Massimiliano, Maurizio Sibilla, and Lamberto Tronchin. 2021. "Energy Modelling and Analytics in the Built Environment—A Review of Their Role for Energy Transitions in the Construction Sector" Energies 14, no. 3: 679. https://doi.org/10.3390/en14030679

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop