Green Energy and Sustainability ISSN 2771-1641

Green Energy and Sustainability 2022;2(3):0006 | https://doi.org/10.47248/ges2202030006

Review Open Access

Evaluation of methods for determining energy flexibility of buildings

Georgios Chantzis 1 , Panagiota Antoniadou 1 , Maria Symeonidou 1 , Elli Kyriaki 1 , Effrosyni Giama 1 , Symeon Oxyzidis 1 , Dionysia Kolokotsa 2 , Agis M. Papadopoulos 1

  • Process Equipment Design Laboratory, Department of Mechanical Engineering, Aristotle University of Thessaloniki, GR 54124 Thessaloniki, Greece
  • Laboratory of Energy Management in the Built Environment, Division of Environmental Management, Technical University of Crete, GR 73100 Chania, Crete, Greece

Correspondence: Georgios Chantzis

Academic Editor(s): Thomas Kotsopoulos, Georgios Martinopoulos, Giorgos Panaras

Received: Nov 4, 2021 | Accepted: Feb 21, 2022 | Published: Jul 4, 2022

This article belongs to the Special Issue

© 2022 by the author(s). This is an Open Access article distributed under the terms of the Creative Commons License Attribution 4.0 International (CC BY 4.0), which permits unrestricted use, distribution, and reproduction in any medium or format, provided the original work is correctly credited.

Cite this article: Chantzis G, Antoniadou P, Symeonidou M, Kyriaki E, Giama E, Oxyzidis S, Kolokotsa D, Papadopoulos A. Evaluation of methods for determining energy flexibility of buildings. Green Energy Sustain 2022; 2(3):0006. https://doi.org/10.47248/ges2202030006

Abstract

The high rate of penetration of renewable energy sources leads to challenges in planning and controlling the production, transmission and distribution of energy. A possible solution lies within the change from traditional supply side management to demand side management. Buildings are good candidates for implementing a demand response model since they account for around 39% of global final energy use and are stably connected to all infrastructure networks. As a result, employing buildings as "players" in energy networks is considered now more than ever compelling. Recently, significant improvement has been denoted in the thermal efficiency of the building shell and the energy efficiency of the HVAC systems in new and renovated buildings. However, despite the reduction in energy demand regarding the space conditioning, buildings continue to be passive end users of the energy system. In order to ensure that they are capable of providing the necessary energy flexibility to balance intermittent energy production, a first step is to establish a formal, standard, and robust method of characterizing the energy flexibility provided on the demand side. Buildings can supply flexibility in a variety of ways, but there is currently no fixed and consistent method for quantifying the amount of flexibility a building can provide to future energy systems. In this paper, an overview of the literature on building energy flexibility will be offered, as well as an introduction to the concept of building energy flexibility and the methodologies used to define and evaluate it.

Keywords

Interactive buildings, Energy Flexibility, Demand Side Management (DSM), Building energy efficiency, Key Performance Indicators (KPIs)

Introduction

Global power usage has risen in recent decades in tandem with rising consumer demand for higher living standards [1]. This increase in global energy demand, combined with forecasts for reduced fossil fuel availability and indications for increased global warming effects have led to increased research interest in RES. Due to the inherent variability that defines energy sources such as wind and solar, the risk of energy network stability issues increases as RES become more dominant in meeting demand [2].

High generation and grid reinforcement costs, inefficient operation, and balancing and reliability issues are all rising concerns for electrical power systems [1]. To address these issues, significant expenditures will be required in the deployment of additional power generating units as well as the reinforcement of transmission and distribution networks [3].

The increased use of renewable energy sources has resulted in a rise in production at the distribution network level, while the entry and active engagement of consumers in the energy market necessitates a reassessment of how DSOs and TSOs operate and collaborate [4].

Traditionally, energy is centrally generated by generation units and delivered through transmission and distribution networks to customers in the classic energy grid paradigm [5]. In this situation, the energy flow is just in one direction. In today's energy networks, the penetration of renewable energy sources and their direct link to the distribution network has added the two-way flow of energy and information (Figure 1) [5].

Figure 1 The flow of energy, services and information in future energy networks [5].

The demand for network stability reinforces the need to manage demand and to ensure flexible energy consumption. Buildings are supposed to play a significant role in this transformation by providing the energy system with their flexibility potential, either in the form of individual buildings or building clusters, to assist in fulfilling the demands of energy networks [2].

While buildings may offer flexibility in a number of ways, there is presently no standard technique for measuring the amount of flexibility they can contribute to future energy systems. Several review papers on building energy flexibility performance indicators have been published.

A study conducted by Reynders et al. 2018 [6], reviewed in several buildings, energy flexibility quantification methods emphasizing solely on thermal energy storage systems in residential buildings. Chen et al. 2018 [7], in their article, classified flexibility metrics and examined the methods for quantifying flexibility of building energy systems, but their review did not examine methods that consider the whole building’s energy flexibility. Lund et al. 2015 [8] concentrated on the potential for supply and demand flexibility of different technologies in their research, but without providing clear indicators for assessing energy flexibility. A short review of energy flexibility quantification methods referring mainly to thermal storage and building appliances was presented by Lopes et al. 2016 [9]. Pean et al. 2019 [10] examined the application of control methods on heat pump systems for enhancing buildings energy flexibility. The acquired flexibility is evaluated using a variety of KPIs presented in literature. Vigna et al. 2018 [11] studied energy flexibility quantitative indicators and methods at the spatial scale of building clusters, while Kathirgamanathan et al. 2021, in their study, [12] examined the flexibility potential of the application of data-driven predictive control methods on buildings with minor focus on flexibility quantification metrics.

Current review papers cover a wide range of metrics, but none of them focuses on the characteristics of energy flexibility that the KPIs or metrics cover. This research gap needs further analysis, as researchers and practitioners in the energy flexibility sector can benefit from a study that focuses on classifying energy flexibility metrics according to a range of respective characteristics. Moreover, in the context of this review another innovation is that the range of established flexibility characteristics is expanded by adding a variety of smart building characteristics that are closely connected to and can help quantify flexibility.

Finally, the rest of the paper consists of four more sections. Section 2 gives an overview of the legislative framework regarding demand side management, buildings and their energy systems and the energy flexibility in future energy networks. Section 3 delves more into the topic of building energy flexibility, including an overview of energy flexibility definitions, building flexibility resources, control strategies, demand-side programs and load shaping methods. The literature review of energy flexibility characteristics and assessment and classification of metrics are introduced in Section 4, while Section 5 focuses on the conclusions, the present study constraints, and the future research possibilities.

Legislative Framework Regarding Buildings, Their Energy Systems and Demand Side Management

The EU wants to lead the transition to clean energy [13,14]. In this direction a target of decreasing GHG emissions is set by at least 55% by 2030 [15]. Simultaneously, EU seeks to improve energy efficiency, achieve worldwide leadership in renewable energy [14], and provide customers with reasonable energy supply conditions [13]. In this regard, it aspires to modernize the economy as well as create growth opportunities for all European citizens [13]. In 2018, the EPBD was revised, aiming to encourage smart building technology while improving consumers' involvement in future energy markets [16]. Consumers are anticipated to play a more active role if they have a range of energy suppliers to choose from, access to accurate energy price comparison tools, and the ability to sell self-generated power.

The idea and use of flexibility were regulated by the European Commission in 2017 [17], setting norms and duties for collaboration and data sharing between TSOs and other entities such as "aggregators". Aggregators are defined as "natural or legal persons combining multiple customer loads or generated electricity for sale, purchase or auction in any electricity market" [18]. Consumer’s participation in the energy market was introduced in 2019 through [18,19], the foundations of a free energy market are laid, and the role of "aggregators" as an intermediary link between consumers and the wholesale market is recognized. In addition, the groundwork for the DSO's independence is being prepared. They are also obliged through directive movements calling for buildings to actively participate in the energy market and the right to sell self-produced energy, as long as network congestion is avoided. Furthermore, the legislation provides for the recruitment of demand management and energy storage technologies, as well as any other resource that will allow the transition to the new reality with the fewest feasible changes and/or expansions to the current energy network.

The development of an SRI will allow the evaluation of buildings in order to maximize their energy, as well as their overall, performance, while also ensuring that users' demands are met [16]. Future buildings will be able to adjust their main function in order to fulfill both users’ needs (demand side) and network constraints (supply side). To achieve this aim, buildings should incorporate automations and energy monitoring technologies, as well as be able to provide relevant information to users about the economic benefits of energy conservation and altering the systems’ function so as to meet the network’s needs. The initial efforts to consolidate and promote the SRI index were taken in August 2018, with the release of a first technical research to establish the index's aims and features [20]. In January 2019, attempts were made for a comprehensive definition of the index in a second research based on the results of the first. This research revealed new ways for calculating the indicator and offered the framework that supports their function and describes them. In the same study, initial efforts were made to define the functions of the building systems and the measured parameters for its computation at the building level [21,22]. In December 2021 the SRI platform was launched. The platform further promotes the SRI and other associated best practices. It serves as a debate point for the main aspects of the SRI, as well as an exchange forum for all interested parties and EU countries [23].

The Concept of Energy Flexibility

Energy flexibility definitions

Various definitions of energy flexibility can be found in the international literature. A general definition describes flexibility as the ability to deviate from design load that characterizes a building or system [24]. In the context of Annex 67 of the IEA, energy flexibility is defined as “the ability of a building to manage its demand and production according to local climatic conditions, user needs and network requirements” [2,25].

In general, flexibility potential is mainly used to reduce energy costs or the cost of purchasing electricity from the grid. This is typically achieved by developing the ability to balance production and demand in real time, in order to maintain the stability of a network even in cases of increased penetration of renewable energy sources [26].

Eventually, energy flexibility can be defined as the ability of a building to react to some external stimulus, expressed as power or energy that can be shifted without compromising the comfort of the interior environment of a building [27]. Finding a standard way to correctly describe the form of those energy or/and power shifts is the main challenge while characterizing flexibility [27]. Flexibility aims at balancing production and demand. Therefore, when describing the power or energy shifting potential, the various objectives and constraints but also the optimal control strategies must be taken into account [28]. In order to achieve and evaluate the energy flexibility of a building a series of steps need to be applied. Firstly, the sources of flexibility are identified, then the building’s loads are divided into two main categories, more “flexible” and less "flexible", and the appropriate control strategies that will allow the flexible operation of systems are selected. Finally, the key performance indicators that will allow the optimal evaluation of energy flexibility are defined and classified [29].

Energy resources of buildings

Energy flexibility in a building is typically achieved by disconnecting the power demand from the power supply, using some form of storage, or smart management of the operation of heating, cooling, air-conditioning, and lighting systems, shifting energy use from periods of high energy cost or network failure to periods of low energy cost [2]. Another approach used, is cutting off the least important loads, without having to restore them later [25]. According to Annex 67 the basic resources of energy flexibility of a building, are [2] the building thermal mass; the thermal energy storage; the fuel switch; the electricity storage; the local electricity production and the connection to energy grids.

The thermal mass of a building is used as a mean of short-term storage (from a few hours to a few days) of thermal energy and is utilized with the aim of load shifting when energy deficiency is observed [30]. The thermal mass includes the whole heat capacity of both the shell and the contents of a building (e.g., walls, floors, furniture, etc.). When utilizing thermal mass to achieve flexibility, special attention must be given to maintaining high thermal comfort levels within the building [2]. Achieving high comfort conditions within the build environment is one of the biggest challenges set not only by the legislation framework but also by the policy makers. Despite the intense research interest towards this direction, limitations and directives regarding the occupants’ comfort and well-being are not yet clearly defined [31–33].

The storage of thermal energy is usually carried out through the storage of hot water for heating or/and DHW in water tanks [2]. Alternatively, PCM can be used as a storage medium instead of water [34]. The use of stored energy will enable load shifting through deactivation of the heating system or postponing DHW production for a certain time period.

Electricity storage is used mainly while applying load shifting strategies. In periods of low load demand the surplus electricity is stored, usually in the form of chemical energy (batteries) and is used in periods where insufficiency of load coverage is observed [35].

Fuel switch refers to the exploitation of the flexibility provided by the provision of energy services using different fuel each time depending on its price and availability. It applies to building systems with more than one power source installed [2].

Electricity generation concerns buildings with integrated systems for local energy production. Usually two types of technologies are considered, renewable energy systems (photovoltaic systems, micro-wind turbines) and small-scale power generation units (Cogeneration/trigeneration units) [36]. The RES systems provide flexibility by covering a portion of the building load, while the cogeneration and trigeneration units, which can offer sufficient controllability and minimal inherent variability, are utilized to maintain network equilibrium [36].

The building becomes more “energy flexible", but also "more resilient", through its connection to more than one energy networks (e.g., electricity grid, gas network, district heating network) [2]. The term “resilient” refers mostly to the thermal resilience of a building [37], except in the case of a building equipped with a gas CHP system, which can utilize the supply of natural gas for electricity generation and therefore becomes “energy resilient” [38].

Control strategies

It is important to adopt a type of control strategy in building energy systems in order to use a building and its systems to offer energy flexibility for the advantage of both customers and utilities [39]. The two major categories in which control strategies are classified, are RBC and MPC [10]. RBC is a basic heuristic technique that monitors the status of a parameter, for example temperature, and sets value limitations for it. When the limits are exceeded then the system responds changing its function according to the predetermined strategy [10,39]. MPC is a more complex method, which bases its function on modelling a building and forecasting its energy behavior [39]. In this case, the most efficient energy management strategy, usually results from solving an optimization problem with some constraints and a specific time horizon [10]. The implementation of any of the above control strategies, requires the existence of controllers installed in the energy system that we wish to control. The output of the control strategy is used as input of the controller. The final control is done through temperature or power regulation or by changing the operating profile of the system [10].

Demand Side Management programs

IBP and PBP are the two basic types of DSM programs [40,7]. Classic and market-based programs are the two types of IBPs. Consumers typically receive rewards in the form of points or a discount for participating in traditional programs, whereas participation in market programs typically leads to a cash reward, based on the amount of demand reduced during periods of high demand and/or reduced production (critical periods) [40,7]. PBPs apply to energy markets in which electricity price is not constant but varies depending on the wholesale price of electricity (dynamic pricing). Within the framework of PBPs application, high or low prices are provided during periods of high or low, respectively, while aiming on balancing production and demand [40,7].

Technical application of demand response

The technical implementation of demand response is accomplished through the use of two forms of control: direct control and indirect control [40,41]. Indirect control involves “aggregators” or utilities transmitting information on energy costs to consumers, who then decide whether or not to adjust their loads to save money. As communication is just an approach in this scenario, the aggregator cannot directly influence load demand; instead, it simply provides the necessary information and monitors any changes in load [40,41]. In the context of direct control, the "aggregator" or utility, whether it has the ability to request an increase or decrease in load during specific time periods or has direct control over individual loads. Consumers inform the "aggregators” about their decisions to shift their load so that they can estimate load demand and network status at a later time during direct control [40,41].

Load shaping methods

Peak clipping, increasing demand or valley filling, shifting load, strategic conservation, general load increase, and flexible load shape are the six fundamental strategies of load shifting (Figure 2) met in the international literature [42]. Peak clipping and valley filling are considered direct load management strategies, whereas load shifting is a mix of the two. Peak clipping is a reasonably simple and highly effective method of reducing peak power demands on a system, usually by direct control of customer loads via signals directed to consumer appliances [43]. Valley filling is another technique that focuses on decreasing the difference between maximum and minimum power demands [44]. During the implementation of valley filling the main objective is increasing demand during off peak hours which is usually accomplished by incentivizing customers to boost their demand [44]. Load shifting assumes the presence of regulated loads and is regarded as an excellent approach for preserving total energy balance as well as user comfort and well-being when each load is cut off and restored at a later time. Strategic conservation reduces overall seasonal energy consumption, mostly by eliminating wasted energy, and so improves system efficiency [45]. On the other hand, strategic load growth leads to an overall increase of seasonal energy consumption [45]. During a flexible load shape customers can purchase some power at lower than usual reliability levels. Depending on the real-time reliability conditions, the customer’s load shape will be flexible [46]. Because they include modifications to the whole shape of a load demand curve [42], the last three approaches need the use of intelligent control algorithms and are categorized as top-level control methods [47].

Figure 2 Load shaping methods [48].

Evaluation of Energy Flexibility Provided

Energy flexibility features

Time, capacity, and efficiency are the three major factors to consider when defining energy flexibility [2,49]. However, flexibility is influenced by a variety of factors (e.g., external environment conditions, building condition, user behavior, etc.), therefore more characteristics must be evaluated to draw safe conclusions about the potential flexibility of a building and its energy systems. Especially, flexibility should be studied as a dynamic phenomenon since one of its main characteristics is its changing nature [50]. The fundamental objective of energy flexibility is to support the seamless integration of renewable energy sources into energy networks. It can also help in improving network stability, lowering energy and CO2 costs, and improving energy management at the building’s level, as well as at the community level (microgrid) [51]. The indicators and methodologies to be utilized must generate meaningful information on energy flexibility both on the demand and production side. It is also critical to have well defined indications and procedures for assessing energy flexibility in the future energy stability system. A distinct context first utilized by Angelakoglou et al. (2019) to evaluate the performance of smart buildings, covers some more elements of energy flexibility and can help achieve the goal of accurately identifying and classifying flexibility indicators [52]. The application of this context will enable the assessment of flexibility as well as the overall performance from a more holistic perspective [29,52]. The main characteristics that can be used to classify flexibility indicators are, technical performance (e.g., energy consumption, load displacement potential, efficiency of on-site RES); environmental performance (e.g., reduction of CO2 emissions); cost-effectiveness (e.g., reducing energy costs) and social performance (e.g., user comfort).

Indicators and methodologies for evaluating energy flexibility

Some indicators and techniques for assessing the potential of energy flexibility of buildings were denoted from the conducted in depth literature analysis based on research from the previous decade. Table 1 presents the indicators and methodologies identified in the literature, as well as the equations used to calculate them, the units of measurement, and the source from which they were obtained. The table's last column contains a brief description of the indicators and the methodology used, as well as the systems and buildings to which they were applied, but also comments on the outcomes of their application and the indicators' adequacy.

Table 1 Energy flexibility indicators denoted in literature.

The metrics are organized into categories based on the characteristics of energy flexibility that they address (Table 2) [29]. Indicators relating to one or two of the aforementioned sources of flexibility are found in the majority of research. In certain publications, indicators covering most elements of flexibility have been presented, but their robustness, quality, and accuracy of the data obtained during their use have not been confirmed [29]. As indicated in Table 2, most of the metrics analyzed are concerned with technical and economic performance, but a few are concerned with social performance, particularly the effects on the internal environment comfort. The technical performance indicators presented in [55] and [62] aim to describe the degree of utilization of on-site energy generation in relation to local energy demand, and can be referred to as load match indicators and PEgrid [68] shows the interaction with the distribution grid and can be considered a grid interaction indicator. Time, in the form of duration and capacity is investigated in fewer research studies, whereas environmental performance is studied in only one case. Moreover, only two surveys namely [54] and [69] presented a Holistic Approach Index, and the index was defined as a weighted average of individual flexibility indicators.

Table 2 Classification of energy flexibility indicators based on their characteristics.

Conclusions

For the adoption of high RES penetration in energy systems, optimizing the balance between production and demand is deemed essential. Buildings have this essential capacity and are capable of leading the clean energy transition. The measurement of the energy flexibility that buildings can provide is in that sense a prerequisite. As a general feature, when a structure can shift large amounts of energy over a long period of time, it is considered more flexible than a structure that can alter smaller amounts of energy over a shorter length of time [27].

According to the findings, there is no commonly accepted method for assessing energy flexibility. Furthermore, it is observed that there is no clear indicator that adequately characterizes it, taking into account all of the influencing factors.

The literature review conducted as part of this study revealed the presence of several indicators for assessing energy flexibility. Direct and indirect indicators are the two types of indicators that may be more easily used in this respect. Indirect indicators include those that measure the building's interaction with the grid, load balancing, energy efficiency, capacity, and thermal comfort of the internal environment, but do not directly assess energy flexibility. Furthermore, demand side management, the capacity to modify load, and the ability to absorb locally produced energy from the building are all tools that, when utilized together, may provide an integrated framework for evaluation.

The indicators were divided into groups based on the characteristics to which they may be related to. The categorization revealed that technical as well as economic and, to a lesser extent, social performance, are given high priority while measuring flexibility.

A variety of problems in recognizing and evaluating the energy flexibility of buildings have been highlighted, as a result of the current restrictions, which constitute topics of interest for future studies. When evaluating the flexibility potential of a building, it became evident that most of the studies do not take into account the flexibility of all building’s loads, but rather focus on thermal or on electrical ones. As a result, the first issue is to establish a specialized technique of modeling a building's flexibility, as it is currently essential to develop numerous models that analyze a building's entire flexibility. A second issue is determining the right criteria along with their weights, which will allow the development of a holistic approach indicators that offer a reliable profile of the available flexibility degree. The integration of quantified user comfort in the flexibility evaluation indicators is the final and most critical issue, as there are currently no indications in case of the influence of flexibility activities regarding comfort conditions in the indoor environment.

In this context, it should be emphasized that the integrated evaluation and realization of the goal of nearly zero energy buildings constitutes a multidimensional problem, rather than a one-dimensional requirement. In particular, all the bibliographic analyses indicate that evaluating the indoor environment conditions while taking into account users’ comfort is needed to properly monitor the construction's operation, thermal comfort conditions, and management of the buildings’ capacity.

Declarations

Ethics Statement

Not applicable.

Consent for Publication

Not applicable.

Availability of Data and Material

Not applicable.

Funding

The authors would like to express their gratitude to the Hellenic Foundation for Research and Innovation, as the research work was supported by the Hellenic Foundation for Research and Innovation (H.F.R.I.) under the “First Call for H.F.R.I. Research Project to support Faculty members and Researchers and the procurement of high-cost research equipment grant” (Project Number: 4104).

Competing Interests

The authors have declared that no competing interests exist.

Author Contributions

All authors contributed to this research. In specific, Georgios Chantzis conceived, designed and performed the review, analyzed the data and wrote the paper; Panagiota Antoniadou analyzed data, contributed to writing and conceptualization, and edited the manuscript; Maria Symeonidou and Elli Kyriaki contributed to writing the manuscript and data analysis; Effrosyni Giama, Symeon Oxyzidis and Dionysia Kolokotsa supervised the manuscript and contributed to conceptualization and validation of the manuscript; Agis M. Papadopoulos contributed to the data analysis, conceptualization and validation, supervised and critical edited the manuscript.

References

1. Heylen E, Deconinck G, Van Hertem D. Review and classification of reliability indicators for power systems with a high share of renewable energy sources. Renew Sustain Energy Rev 2018;97:554-568. [Google Scholar] [CrossRef]
2. Jensen SØ, Marszal-Pomianowska A, Lollini R, Pasut W, Knotzer A, Engelmann P, et al. IEA EBC Annex 67 Energy Flexible Buildings. Energy Build 2017 Nov 15;155:25-34. [Google Scholar] [CrossRef]
3. Jordehi AR. Optimisation of demand response in electric power systems, a review. Renew Sustain Energy Rev 2019 Apr 1;103:308-319. [Google Scholar] [CrossRef]
4. ENTSO-E. Towards smarter grids : Developing TSO and DSO roles and interactions for the benefit of consumers. Eur Netw Transm Syst Oper Electr 2015;:1-8. [Google Scholar]
5. IRENA. Co-Operation Between Transmission and Distribution System Operators Innovation Landscape Brief About Irena [Internet]. [cited 2021 Sep 23] Available from: www.irena.org.
6. Reynders G, Amaral Lopes R, 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]
7. Chen Y, Xu P, Gu J, Schmidt F, Li W. Measures to improve energy demand flexibility in buildings for demand response (DR): A review. Energy Build 2018;177:125-139. [Google Scholar] [CrossRef]
8. Lund PD, 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]
9. Lopes RA, Chambel A, Neves J, Aelenei D, Martins J. A Literature Review of Methodologies Used to Assess the Energy Flexibility of Buildings. Energy Procedia 2016 Jun 1;91:1053-1058. [Google Scholar] [CrossRef]
10. Péan TQ, 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 Feb 1;74:35-49. [Google Scholar] [CrossRef]
11. Vigna I, Pernetti R, Pasut W, Lollini R. Literature review on energy flexibility definitions and indicators for building clusters. [Internet]. [cited 2021 Oct 22] Available from: https://annex67.org/media/1863/literature-review-on-energy-flexibility-definitions-and-indicators-for-building-clusters.pdf.
12. Kathirgamanathan A, De Rosa M, Mangina E, Finn DP. Data-driven predictive control for unlocking building energy flexibility: A review. Renew Sustain Energy Rev 2021 Jan 1;135:110120. [Google Scholar] [CrossRef]
13. European Parliament; L 315. Directive 2012/27/EU of the European Parliament and of the Council of 25 October 2012 on energy efficiency. Off J Eur Union [Internet]. [cited 2021 Oct 2] Available from: http://data.europa.eu/eli/dir/2012/27/oj.
14. European Parliament. Directive (EU) 2018/2001 of the European Parliament and of the Council on the promotion of the use of energy from renewable sources. Off J Eur Union [Internet] 2018;L 328:82-209. (December 2018). [Google Scholar]
15. European Commission. Regulation (EU) 2021/1119 of the european parliament and of the council establishing the framework for achieving climate neutrality and amending Regulations (EC) No 401/2009 and (EU) 2018/1999 (‘European Climate Law’). Off J Eur Union [Internet] 2021;2021:17. (June). [Google Scholar]
16. European Parliament. Directive 2018/2002/EU of the European Parliament and of the Council of 11 December 2018 amending Directive 2012/27/EU on Energy Efficiency [Internet]. Vol. L 328, Official Journal of the European Union. [cited 2021 Oct 2] Available from: http://data.europa.eu/eli/dir/2018/2002/oj.
17. European Commission. Commission Regulation (EU) 2017/1485: Establishing a guideline on Electricity Transmission System Operation (SO GL). Off J Eur Union [Internet] 2017;L 220:120. (25 August 2017). [Google Scholar]
18. European Parliament. Directive 2019/944 on Common Rules for the Internal Market for Electricity. Off J Eur Union. [Internet]. (L 158/125).
19. European Parliament. Regulation (EU) 2019/943 of the European Parliament and of the Council of 5 June 2019 on the internal market for electricity. Off J Eur Union. [Internet] 2019;62(L158):54-191. [cited 2021 Oct 8] Available from: http://data.europa.eu/eli/reg/2019/943/oj. [Google Scholar]
20. Tichelen P Van, Bogaert S. Support for Setting Up a Smart Readiness Indicator for Buildings and Related Impact Assessment Final Report 2018. (August).
21. Verbeke S, Aerts D, Rynders G, Yixiao M. Interim report July 2019 of the 2th technical support study on the Smart Readiness Indicator for Buildings. VITO Rep 2019;:281. (JULY). [Google Scholar]
22. Verbeke S, Aerts D, Reynders G, Ma Y, Waide P. Final Report on the Technical Support To the Development of a Smart Readiness Indicator for Buildings [Internet]. Available from: https://op.europa.eu/en/publication-detail/-/publication/f9e6d89d-fbb1-11ea-b44f-01aa75ed71a1/language-en.
23. Smart readiness indicator | Energy [Internet]. [cited 2021 Dec 24]. Available from: https://ec.europa.eu/energy/topics/energy-efficiency/energy-efficient-buildings/smart-readiness-indicator_en.
24. Petersen MK, Edlund K, Hansen LH, Bendtsen J, Stoustrup J. A taxonomy for modeling flexibility and a computationally efficient algorithm for dispatch in Smart Grids. Proc Am Control Conf 2013;:1150-1156. [Google Scholar] [CrossRef]
25. Marszal AJ, Pernetti Roberta I, Søren Østergaard Jensen D, Marszal AJ, Pernetti Roberta I, Søren Østergaard Jensen D. Characterization of Energy Flexibility in Buildings Energy in Buildings. IEA EBC Annex 67 report 2019. (December). [Google Scholar]
26. Clauß J, Stinner S, Solli C, Lindberg KB, Madsen H, Georges L. A generic methodology to evaluate hourly average CO2eq. intensities of the electricity mix to deploy the energy flexibility potential of Norwegian buildings The future Norwegian Energy system in a European context View project IEA EBC Annex 67 Energy Flex. [Internet].
27. D’hulst R, Labeeuw W, Beusen B, Claessens S, Deconinck G, Vanthournout K. Demand response flexibility and flexibility potential of residential smart appliances: Experiences from large pilot test in Belgium. Appl Energy 2015 Oct 1;155:79-90. [Google Scholar] [CrossRef]
28. Eid C, Codani P, Chen Y, Perez Y, Hakvoort R. Aggregation of demand side flexibility in a smart grid: A review for European market design. Int Conf Eur Energy Mark EEM 2015;2015-Augus:10-14. (May). [Google Scholar] [CrossRef]
29. Chantzis G, Antoniadou P, Symeonidou M, Giama E, Oxyzidis S, Kolokotsa D, et al. 12th National Conference On Renewable Energy Sources (Greek); 2021. Thessaloniki: Solar Institute; 2021 Evaluation of methods for determining Energy Flexibility of Buildings.. p.763-774. In:.
30. Lu F, Yu Z, Zou Y, Yang X. Cooling system energy flexibility of a nearly zero-energy office building using building thermal mass: Potential evaluation and parametric analysis. Energy Build 2021;236:110763. [Google Scholar] [CrossRef]
31. Antoniadou P, Papadopoulos AM. Occupants’ thermal comfort: State of the art and the prospects of personalized assessment in office buildings. Energy Build 2017;153:136-149. [Google Scholar] [CrossRef]
32. Antoniadou P, Papakostas KT, Papadopoulos AM. The occupants’ comfort in non-residential nearly Zero Energy Buildings in the 21st century: A review. [CrossRef]
33. Antoniadou P, Chantzis G, Symeonidou M, Giama E, Oxyzidis S, Kolokotsa D, et al. 12th National Conference On Renewable Energy Sources (Greek); 2021. Thessaloniki: Solar Institute; 2021 Evolution of thermal comfort requirements in the building sector.. p.443-51. In:.
34. Arteconi A, Xu J, Ciarrocchi E, Paciello L, Comodi G, Polonara F, et al. Demand Side Management of a Building Summer Cooling Load by Means of a Thermal Energy Storage. Energy Procedia 2015;75:3277-3283. [Google Scholar] [CrossRef]
35. Groppi D, Pfeifer A, Garcia DA, Krajačić G, Duić N. A review on energy storage and demand side management solutions in smart energy islands. Renew Sustain Energy Rev 2021 Jan 1;135:110183. [Google Scholar] [CrossRef]
36. Tang H, Wang S, Li H. Flexibility categorization, sources, capabilities and technologies for energy-flexible and grid-responsive buildings: State-of-the-art and future perspective. Energy 2021;219:119598. [Google Scholar] [CrossRef]
37. Homaei S, Hamdy M. Thermal resilient buildings: How to be quantified? A novel benchmarking framework and labelling metric. Build Environ 2021 Aug 15;201:108022. [Google Scholar] [CrossRef]
38. EPA. Guide to Using Combined Heat and Power for Enhancing Reliability and Resiliency in Buildings [Internet]. [cited 2021 Dec 24] Available from: https://www.epa.gov/sites/production/files/2015-07/documents/guide_to_using_combined_heat_and_power_for_enhancing_reliability_and_resiliency_in_buildings.pdf.
39. Clauß J, Finck C, Vogler-finck P, Beagon P. Control strategies for building energy systems to unlock demand side flexibility – A review Norwegian University of Science and Technology, Trondheim, Norway Eindhoven University of Technology, Eindhoven, Netherlands Neogrid Technologies ApS / Aalborg 2017;:611-620. (May 2018).
40. Albadi MH, El-Saadany EF. A summary of demand response in electricity markets. Electr Power Syst Res 2008;78(11):1989-1996. [Google Scholar] [CrossRef]
41. Madsen H, Parvizi J, Halvgaard R, Sokoler LE, Jørgensen JB, Hansen LH, et al. In: Handbook of Clean Energy Systems. Chichester, UK: John Wiley & Sons, Ltd; 2015. p.1-26. [Google Scholar]
42. Lokeshgupta B, Sadhukhan A, Sivasubramani S. Multi-objective optimization for demand side management in a smart grid environment. [cited 2021 Sept 13] Available from: https://ieeexplore.ieee.org/document/8387293/.
43. Gellings CW. Evolving practice of demand-side management. J Mod Power Syst Clean Energy 2017;5(1):1-9. [Google Scholar] [CrossRef]
44. Attia HA. Mathematical Formulation of the Demand Side Management (DSM) Problem and its Optimal Solution 2010;10:953-959.
45. Sa’ed JA, Wari Z, Abughazaleh F, Dawud J, Favuzza S, Zizzo G. Effect of demand side management on the operation of pv-integrated distribution systems. Appl Sci 2020;10(21):1-26. [Google Scholar] [CrossRef]
46. Luo T, Ault G, Galloway S. Demand side management in a highly decentralized energy future. Proc Univ Power Eng Conf 2010. [Google Scholar]
47. Logenthiran T, Srinivasan D, Shun TZ. Demand side management in smart grid using heuristic optimization. IEEE Trans Smart Grid 2012;3(3):1244-1252. [Google Scholar] [CrossRef]
48. Jabir HJ, Teh J, Ishak D, Abunima H. Impact of Demand-Side Management on the Reliability of Generation Systems. Energies 2018;11:2155. [Google Scholar] [CrossRef]
49. Reynders G, Diriken J, Saelens D. Generic characterization method for energy flexibility: Applied to structural thermal storage in residential buildings. Appl Energy 2017;198:192-202. [Google Scholar] [CrossRef]
50. Junker RG, Azar AG, Lopes RA, Lindberg KB, Reynders G, Relan R, et al. Characterizing the energy flexibility of buildings and districts. Appl Energy 2018 Sep 1;225:175-182. [Google Scholar] [CrossRef]
51. Shayeghi H, Shahryari E, Moradzadeh M, Siano P. A survey on microgrid energy management considering flexible energy sources. Energies 2019;12(11):1-26. [Google Scholar] [CrossRef]
52. Angelakoglou K, Nikolopoulos N, Giourka P, Svensson I-L, Tsarchopoulos P, Tryferidis A, et al. A Methodological Framework for the Selection of Key Performance Indicators to Assess Smart City Solutions. Smart Cities 2019;2(2):269-306. [Google Scholar] [CrossRef]
53. Nuytten T, Claessens B, Paredis K, Van Bael J, Six D. Flexibility of a combined heat and power system with thermal energy storage for district heating. Appl Energy 2013;104:583-591. [Google Scholar] [CrossRef]
54. Arteconi A, Mugnini A, Polonara F. Energy flexible buildings: A methodology for rating the flexibility performance of buildings with electric heating and cooling systems. Appl Energy 2019 Oct 1;:251. [Google Scholar] [CrossRef]
55. Reynders G, Nuytten T, Saelens D. Potential of structural thermal mass for demand-side management in dwellings. Build Environ 2013 Jun;64:187-199. [Google Scholar] [CrossRef]
56. De Coninck R, Helsen L. Quantification of flexibility in buildings by cost curves - Methodology and application. Appl Energy 2016 Jan 15;162:653-665. [Google Scholar] [CrossRef]
57. Oldewurtel F, Sturzenegger D, Andersson G, Morari M, Smith RS. Proceedings of 52nd IEEE Conference on Decision and Control; 2013 Dec 10-13;. Towards a Standardized Building Assessment for Demand Response.. [CrossRef]
58. 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(6):800-811. [Google Scholar] [CrossRef]
59. Le Dréau J, Heiselberg P. Energy flexibility of residential buildings using short term heat storage in the thermal mass. Energy 2016 Sep 15;111:991-1002. [Google Scholar] [CrossRef]
60. Taddeo P, Colet A, Carrillo RE, Canals LC, Schubnel B, Stauffer Y, et al. Management and activation of energy flexibility at building and market level: A residential case study. Energies 2020;13(5):1-18. [Google Scholar] [CrossRef]
61. Athienitis AK, Dumont E, De T. Development of a dynamic energy flexibility index for buildings and their interaction with smart grids. 2020 ACEEE Summer Study Energy Effic Build 2020;:31-43. (August). [Google Scholar]
62. Fambri G, Badami M, Tsagkrasoulis D, Katsiki V, Giannakis G, Papanikolaou A. Demand flexibility enabled by virtual energy storage to improve renewable energy penetration. Energies 2020;13(19):1-20. [Google Scholar] [CrossRef]
63. Péan TQ, Torres B, Salom J, Ortiz J. Representation of daily profiles of building energy flexibility. eSim 2018;:153-162. (May). [Google Scholar]
64. D’Ettorre F, De Rosa M, Conti P, Testi D, Finn D. Mapping the energy flexibility potential of single buildings equipped with optimally-controlled heat pump, gas boilers and thermal storage. Sustain Cities Soc 2019 Oct 1;50:101689. [Google Scholar] [CrossRef]
65. Johra H, Heiselberg P, Dréau J Le. Influence of envelope, structural thermal mass and indoor content on the building heating energy flexibility. Energy Build 2019 Jan 15;183:325-339. [Google Scholar] [CrossRef]
66. Yin R, Kara EC, Li Y, DeForest N, Wang K, Yong T, et al. Quantifying flexibility of commercial and residential loads for demand response using setpoint changes. Appl Energy 2016 Sep 1;177:149-164. [Google Scholar] [CrossRef]
67. Tahersima F, Madsen PP, Andersen P. Proceedings of the IEEE International Conference on Control Applications; 2013 Aug 28-30; Hyderabad:India. An intuitive definition of demand flexibility in direct load control..
68. Shen L, Sun Y. Performance comparisons of two system sizing approaches for net zero energy building clusters under uncertainties. Energy Build 2016 Sep 1;127:10-21. [Google Scholar] [CrossRef]
69. De Groote M, Volt J, Bean F. Is Europe ready for the smart buildings revolution?. Mapping smart-readiness and innovative case studies [Internet]. [cited 2021 Feb 13] Available from: www.bpie.eu.
Download PDF

Share this article

About Us Journals Join Us Submit Fees Contact