Next Article in Journal
A Universal Detection Method for Adversarial Examples and Fake Images
Previous Article in Journal
The Influence of a Coherent Annotation and Synthetic Addition of Lung Nodules for Lung Segmentation in CT Scans
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Towards Personalization of Indoor Air Quality: Review of Sensing Requirements and Field Deployments

1
Department of the Built Environment, College of Design and Engineering, National University of Singapore, Singapore 119077, Singapore
2
Department of Construction Science and Management, Clemson University, Clemson, SC 29634, USA
3
School of Engineering and Applied Sciences, Western Kentucky University, Bowling Green, KY 42101, USA
4
Trane Technologies PLC Engineering & Technology Centre, Bangalore 560029, India
*
Author to whom correspondence should be addressed.
Sensors 2022, 22(9), 3444; https://doi.org/10.3390/s22093444
Submission received: 31 March 2022 / Revised: 20 April 2022 / Accepted: 26 April 2022 / Published: 30 April 2022
(This article belongs to the Section Environmental Sensing)

Abstract

:
As humans spend more time indoors, ensuring acceptable indoor air quality (IAQ) through ubiquitous sensing systems has become a necessity. Although extensive studies have been conducted on the IAQ sensing systems, a holistic review of the performance and deployment of Ubiquitous IAQ Sensing (UIAQS) systems with associated requirements in IAQ sensing standards is still lacking. In this study, we first reviewed IAQ pollutants and other IAQ-related factors and the associated requirements in the prominent IAQ sensing standards. We found that while non-pollutant factors are influential on occupants’ perception of IAQ and their satisfaction, they do not have evaluation metrics in the IAQ standards. Then, we systematically reviewed field studies on UIAQS technologies in the literature. Specific classes of information were recorded and analyzed further. We found that the majority of the UIAQS systems did not meet the requirements of the prominent IAQ sensing standards and identified four primary research gaps. We concluded that a new holistic and personalized approach that incorporates UIAQS measurements and subjective feedback is needed. This study provides valuable insights for researchers and policymakers to better improve UIAQS technologies by developing personalized IAQ sensors and sensing standards.

1. Introduction

Indoor Environmental Quality (IEQ) has gained increasing importance since humans spend the majority (up to 90%) of their time indoors [1]. Among factors influencing Indoor Environmental Quality (IEQ), Indoor Air Quality (IAQ) is becoming more prominent as our state of knowledge on the role of air in the spread of pathogens, pollens, and chemical substances advances. These pollutants significantly contribute to Sick Building Syndrome (SBS), which is defined as discomfort or acute health symptoms experienced by occupants in indoor environments [2]. In addition, poor indoor air quality is known to negatively affect occupant productivity [3]. To address the challenges of energy efficiency in buildings and improve occupant comfort, health, and productivity, IAQ standards and codes, such as ASHRAE 62.1, World Health Organization (WHO) 2006, and Singapore Standard SS 554, provide guidelines on IAQ pollutant limits [4].
It is very challenging to monitor all IAQ pollutants with high accuracy and resolution (e.g., accurate measurements at the occupant level) at scale. In addition, other IAQ-related factors, such as odor and humidity that may not directly have adverse impacts on human health, yet may impact on occupants’ satisfaction and comfort, are often not considered via quantifiable metrics in IAQ standards [5]. For instance, although odors may not be directly harmful to human health, they play an essential role in influencing occupants’ mood, alertness, and cognitive performance [6]. Furthermore, studies have shown that humidity can affect the way people perceive IAQ, and it could impact occupants’ emotions [7]. In the absence of sensing methods integrated with IAQ perception, occupants take control of air ventilation systems (e.g., natural ventilation, HVAC systems, and fans) based on their own preferences whenever they are offered the opportunity [8]. Another missing category of airborne hazards is microbes that have become a dominant concern in recent years and more particularly with the emergence of the COVID-19 pandemic. The implications of airborne transmission of COVID-19 causes a paradigm shift in the way that people view indoor air pollutants as particulates, potentially carrying the virus and increasing virus incidence [9].
The recent advancements in IAQ sensing technologies and large-scale deployment of IAQ sensing systems and user interfaces (e.g., kiosks and tablets) have provided the opportunity to collect IAQ data (as defined by IAQ standards) and other related factors at high resolution (e.g., the occupant scale). However, few literature reviews are conducted to holistically study IAQ-related factors, including the intersection of IAQ and sensing technologies. For example, H. Zhang and R. Srinivasan [10] reviewed IAQ sensing standards and current commercially available IAQ sensing technologies. However, the existing literature review studies are not holistic. For instance, while Coulby et al. [11] and Saini et al. [12]’s studies focused on the development of IAQ sensing technologies at the building scale, future studies on the IAQ sensing technologies on a larger scale are needed. More specific information, such as deployment locations and duration, are expected in the future design and deployment of IAQ sensing technologies [13]. In addition, there has been no study to delineate the state-of-the-art field deployment of ubiquitous sensing technologies to evaluate their sensing performance and understand their current practices, shortcomings, and future directions, such as the personalization of IAQ.
To address the above-mentioned IAQ challenges, this study first conducts an in-depth review regarding IAQ standards, as well as Ubiquitous IAQ Sensing (UIAQS) systems that could potentially fulfill the standard requirements at the occupant level and in real-time. UIAQS systems can be defined as sensing systems that can provide real-time and a high spatial and temporal resolution for monitoring indoor air concentration levels [14]. To determine the scope of the review, we considered IAQ pollutants and relevant factors impacting occupant perception of IAQ (e.g., odor and humidity) that are neglected in the existing standards. Then, specific standards regarding the requirements of IAQ sensing systems are thoroughly reviewed. This is followed by a systematic review and a comparison of the UIAQS systems deployed in the field and four major categories of research gaps and future directions. By examining the reviewed articles, this study underlines the importance of integrating personal preferences with UIAQS systems in light of a holistic view of the IAQ. Moreover, this study serves as a useful reference for researchers and practitioners to understand current IAQ standards and personalized sensing technologies, which helps establish personalized IAQ sensing (PIAQS) systems.
The structure of the study is as follows. Section 2 provides a systematic review of primary IAQ-related factors, followed by a critical review of current IAQ sensing standards in Section 3. Section 4 discusses the research design of this study. Section 5 presents the systematic review of UIAQS technologies in the field studies. The research gaps and suggestions are discussed in Section 6. The proposed future research direction–developing personalized IAQ sensing systems is elaborated in Section 7, followed by the conclusions in Section 8.

2. Physical Environmental Factors Based on IAQ Standards’ Requirements

In this study, we first provide a short review of the physical environmental factors affecting IAQ based on the IAQ standards requirements, including ASHRAE 62.1, World Health Organization (WHO) 2006, and Singapore Standard SS 554. ASHRAE standards 62.1 is a well-recognized guideline for ventilation system design and indoor air quality, especially where building HAVC systems are used. WHO 2006 air quality standard is widely used in a general indoor environment and the guidelines are now seen as the key source on which the European Commission’s directive on air quality is based [15]. Considering the diversity of climate zones, Singapore Standard SS 554 was also included to represent the IAQ standards adopted in tropical climate regions. The selected physical environmental factors are categorized into four subgroups-IAQ pollutants, odor, humidity, and microbes. Specifically, seven IAQ pollutants, including particulate matter, formaldehyde, carbon dioxide, carbon monoxide, nitrogen dioxide, ozone, and total volatile organic compounds (TVOCs), are selected. Although the temperature is a characteristic of air, we opted not to include it in this review as it is heavily studied within the thermal comfort research domain.

2.1. IAQ Requirements on Air Pollutants

  • Particulate Matter (PM)
Particulate matter is the most commonly monitored indoor air pollutant, which is defined as a mixture of solid or liquid particles suspended in air [16]. These particles vary in size, shape, and composition (e.g., PM1, PM2.5, and PM10,). Specifically, PM10, PM2.5, and PM1 are the particulate matters with aerodynamic diameters less than 10 μm, 2.5 μm and 1 μm [17]. Particulate matter exposure is harmful and dangerous, as it may lead to a variety of adverse health impacts, including eye, nose, and throat irritation, aggravation of coronary and respiratory disease symptoms, and even premature death in people with heart or lung disease [18]. Specifically, J.M. Logue, P.N. Price, M.H. Sherman, and B.C. Singer (2012) [19] reported that PM2.5 has the significant chronic impact on human health, because they are inhalable and not easily filtered by the human lungs, which may cause the particles to enter the bloodstream. More importantly, PM1 may create greater environmental and health issues among other indoor air pollutants, because these smaller-size particles can be inhaled deeper into the lungs and contaminate human body with toxic and harmful substances than the larger particles [20].
Given the adverse effects caused by particulate matter, various agencies provided guidelines on recommended levels of PM10 and PM2.5. Specifically, ASHRAE 62, Singapore Standard SS 554 and WHO 2006 guidelines provided an identical recommended level of PM10, which is 50 μg/m3. WHO 2006 also recommended a limit of PM2.5, which is 25 μg/ m3 [21]. However, PM1 has not been addressed by the majority of the IAQ standards.
  • Formaldehyde
Following the particulate matter, the second most reported indoor pollutant is formaldehyde (HCHO), which was identified as a Group 1 human carcinogen by the World Health Organization (WHO) in 2004 [22]. In indoor air, formaldehyde can be produced by wood-based products and combustion [23]. As the air-exchange rate may become low in residential dwellings where wood is largely adopted, indoor formaldehyde concentrations tend to be generally higher than outdoors. Guidelines for indoor formaldehyde exposure have been established worldwide. For instance, the WHO set the guideline to 100 μg/ m3 in 1987 [24]. Depending on various guidelines, the values of indoor formaldehyde concentration range from 9 μg/m3 to 125 μg/m3 [25].
  • Carbon Dioxide (CO2)
CO2 is another indoor air pollutant that is closely linked to the ventilation rates and occupancy of a space. While occupant respiration is the primary source of CO2, open flame and the use of reinforced concrete could generate significant levels of CO2 [26]. Inadequate ventilation or over-occupancy may increase the concentration of CO2 indoors. At very low air movements, an occupant may inhale from the CO2-rich bubble formed around the breathing zone with significantly higher CO2 concentrations compared to the background [27]. Unhealthy levels of CO2 concentration indoors may have a negative impact on occupant cognitive functions, productivity, and comfort [28]. Inverse correlations between cognitive performance and concentration levels with CO2 have been shown to be the main cause of SBS symptoms [29,30]. A general rule of thumb of the acceptable limit of indoor CO2 level is 700 ppm above outdoor [31], which is aligned with the requirement by Singapore Standard SS 554 in 2021 [32]. While based on ASHRAE, the suggested indoor CO2 level is 1000 ppm in 2019 [33].
  • Carbon Monoxide (CO)
Carbon Monoxide is an odorless and colorless gas that poses potential health hazards to building occupants. Low CO concentrations can cause fatigue in healthy residents and chest pain in people with heart disease [34]. At higher concentrations, it may impair vision and affect coordination and induce headaches, dizziness, confusion or nausea [22]. CO concentration is fatal at very high levels, as carbon monoxide poisoning results in death [35]. The guidelines on the limits of the CO level vary from 6ppm (according to WHO 2006 guidelines) to 9 ppm (according to ASHRAE in 2019 and Singapore Standard SS 554 in 2021) [21,32,33].
  • Nitrogen Dioxide (NO2)
Nitrogen Dioxide is a by-product of combustion processes. Some common sources of NO2 include stoves, ovens, candles, and mosquito coils. Low-level NO2 exposure may increase the risk of respiratory inflammation and infections of building occupants. Extremely high-dose exposure to NO2 may result in pulmonary edema and diffuse lung injury [36]. Continued exposure to high NO2 levels can contribute to the development of acute or chronic bronchitis [22]. According to WHO 2006 guidelines, the suggested limit of nitrogen dioxide in an indoor environment is 200 μg/m3 [21].
  • Total Volatile Organic Compounds (TVOCs)
Total volatile organic compounds (TVOCs) are emitted as gases from certain solids or liquids in the building environment. TVOCs can be found in paint, aerosol sprays such as cleansers, disinfectants, air fresheners, and insecticides. They can also be emitted from building materials like office equipment such as photocopiers and printers, correction fluids and carbonless copy paper, and adhesives like furniture glue. TVOCs can cause acute symptoms such as eye, nose, and throat irritation, headaches, loss of coordination, and nausea. Long-term exposure may even bring damage to the liver, kidney, and central nervous system (e.g., tremors and hepatitis) [37]. Singapore Standard SS 554 suggests that the concentration of TVOCs indoors should be less than 3000 ppb in 2021 [32].
  • Ozone
Ozone is a common air pollutant worldwide, the levels of which have increased globally in recent decades [38]. Ozone can be generated from a chemical reaction between NO2 and TVOCs in exposure to sunlight. The sources of ozone include the emission of chemical solvents, electric utilities, gasoline vapors, and the use of photo-copy devices. It has been widely reported that ozone is associated with respiratory illness, such as cardiovascular mortality [39], stroke [40], and heart failure [41]. People with underlying diseases, children, and the elderly are the highest risk populations for ozone pollutants. Because of the severe consequence of exposure to ozone, the concentrations of ozone in indoor air are regulated by IAQ guidelines. Specifically, the limit of ozone in an indoor environment is 60 ppb based on the guideline of WHO 2006 [22], while the number is 50 ppb according to Singapore Standard SS 554 in 2021 [32].
The summary of the potential health effects and various IAQ code of practice for the selected indoor air pollutants is presented in Table 1.

2.2. Odor

Compared with the IAQ pollutants listed above, odor may not directly threaten occupant health, but it can potentially affect occupant mood, perception of indoor air freshness, productivity and indicate the presence of mold, which negatively impacts occupant’s health by causing irritation, allergic reactions and other respiratory ailments [6]. One of the most important features of odor is that the perception of odor in human brains can be subjective. The same scent can be labeled as “pleasant” or “unpleasant”, depending on an individual’s preferences. To understand subjective indoor air quality, prior studies have provided occupants questionnaires to collect their perceived IAQ [42]. It has been found that personal factors, such as age and gender, affect occupant perception of good indoor air quality. Although humans might adapt to odor over time, which means a person’s perceived IAQ can be initially unpleasant and changes over long exposure durations, no reviewed IAQ standards included the guidelines on the perceived IAQ for control purposes in the built environment.

2.3. Humidity

In addition to common air pollutants, indoor humidity can be categorized as relative humidity (RH) and absolute humidity (AH). Specifically, RH refers to the measure of water vapor present in the air as a percentage of the air vapor capacity (i.e., the amount of water vapor that saturates the air). RH refers to the measure of water vapor relative to the temperature of the air, which is often represented as a measure of humidity indoors. AH refers to the measure of the absolute grains of water vapor in the air, and it is independent of air temperature. Though humidity is not directly an indoor air pollutant, it affects the way occupants perceive IAQ [7]. Current literature supported the claim that RH at an adequate level can positively influence the perception of IAQ [43,44] and sleep quality [45,46]. However, RH less than 30% may lead to the perception of sensory irritants in the air and eye dryness, and RH greater than 70% may cause occupant perception of poor IAQ [47]. Presently, the widely accepted comfortable range of RH is between 30 to 50% [48]. Furthermore, low RH (<30%) can affect human health via two mechanisms. (a) Pathogenic agents’ infectivity and survival have been shown to depend, at least partially, on RH levels [49], and (b) low levels of RH indoors may lead to dry air pathways (i.e., from mouth and nose to the lungs) in the body, which in turn accommodates a likelihood of sickness when one is exposed to infective agents [50].

2.4. Microbes

Poor IAQ can lead to a greater virus incidence beyond the multitude of health problems discussed earlier. Previous studies have shown that air pollution particles have the capacity to carry and spread microbes across distances, increasing airborne diseases (such as SARS, Measles, and Chickenpox) [51]. Long-term exposure to indoor air carrying microbes may weaken human respiratory systems and make residents more susceptible to viruses such as COVID-19 [52]. However, given the potential hazards that microbes may cause, acceptable ranges for various microbes have not been determined by the majority of IAQ standards or guidelines.

3. Requirements of Prominent IAQ Sensing Standards

With the knowledge of the harmful effects and standards for indoor air pollutants, odor, humidity, and microbes, it is vital to develop guidelines and standards for IAQ sensors in assessing air quality. Many organizations, including the European Union (EU), ASTM international, and the US EPA, are considered as standards assessing air sensor performance. Considering data availability and completion, we compared the accuracy requirements of three commonly recognized specification requirements–US EPA [53], WELL (v2) [54], and RESET (v2) [55], which have been illustrated in Figure 1. As shown in Figure 1, no standardized guidelines for IAQ sensing technologies have been developed. The required air characteristics and the corresponding accuracy requirements vary in different standards. Specifically, US EPA has the most completed requirements on the threshold concentration of indoor air pollutants, while only three air pollutants are included in RESET (v2). Compared with another two standards, WELL (v2) has the strictest accuracy requirements on IAQ sensing technologies except for CO. Such different requirements can be attributed to different organizations or countries who were involved in making the standards. Although the measured pollutants vary in different standards, the air pollutants carbon dioxide, TVOCs, and particulate matter are presented across various guidelines, indicating the prevalence of monitoring these indoor air pollutants.

4. Research Methodology

4.1. Selection of Academic Publications

The definition of Ubiquitous Air Quality Sensing (UIAQS) system has merged from several laboratories for industrial practice in recent years, which refers to the sensing systems that can provide real-time and high spatial and temporal resolution for monitoring indoor air concentration levels [14]. This study follows this definition, which provides the key criteria for the selection of academic articles. The selected academic articles were examined based on a four-step review process, which is illustrated in Figure 2.
In step one, the literature database and search rules are defined. To acquire all potential research articles, the literature search was carried out based on the databases ScienceDirect, Scopus, and JSTOR. As this review mainly focuses on the prior studies related to IAQ sensing systems, other databases, such as PubMed, were not included. The search was limited to the last 12 years of publication (i.e., from 2010 to 2021), as UIAQS technology was not prolific before that. In addition, only peer-reviewed papers published in English were included. In this study, “ubiquitous” and “low-cost” are used interchangeably since the expensive sensors are difficult to be applied on large scales. Thus, the keywords “low-cost or inexpensive, indoor air quality, indoor air pollutant, IAQ sensing technologies” are used to narrow down the selection of relevant articles. The second step is to conduct a preliminary search. The primary emphasis of literature selection involved reading titles and abstracts of each paper to exclude the irrelevant articles. For instance, the articles without critical information, such as measured factors, sensor modules, and deployment location, are excluded in this review. After the preliminary literature selection, literature filtration was conducted through reading the remaining articles to ascertain their relevance. In total, 229 papers were identified, and 33 (13%) papers were rigorously reviewed, while the remaining articles were partially reviewed as supporting records, which were referred to only when needed. The final step involved the categorization of the selected literature based on the categorization rules elaborated on in Section 4.2.

4.2. Categorization of UIAQS Technologies

Based on the common features among UIAQS studies, identified IAQ-related factors, and the guidelines for IAQ sensing technologies, this study dissected all selected UIAQS technology-related studies by extracting detailed information on “study location”, “deployment”, “duration”, “measured factors”, “sensor modules”, “measurement range” “accuracy”, “sensor type”, and “occupant information and feedback”. The detailed information of each essential property is presented in Table 2.

5. Findings from Ubiquitous Sensing Technologies in the Field Studies Literature

UIAQS technologies have emerged from different laboratories to practical applications in recent years. Table 3 summarized 33 existing UIAQS studies capable of assessing the concentrations of identified IAQ-related factors.
As shown in Table 3a, when assessing IAQ-related factors, different studies adopted various sensing modules, which may have different measurement ranges and accuracy rates. Given the three IAQ sensing standards reviewed in Section 3, Figure 3 illustrates the percentage of the reviewed sensor modules (with various tested air pollutants) that satisfied or unsatisfied the requirements of sensing standards US EPA, WELL v2, and RESET v2. The majority of the deployed IAQ sensors failed to fully meet the requirement of current commonly adopted IAQ sensing standards. Specifically, none of the sensing modules detecting PM10 satisfy any of the selected IAQ sensing standards. Sensors measuring CO2 have the highest percentage of fulfilling any one of the three IAQ sensing standards. Over 75% of the CO2 sensors meet the requirements of US EPA, while the percentage of CO2 sensors that satisfy the requirements of RESET v2 drops to 38.89%. Only 16.67% of the CO2 sensors meet the requirements of WELL v2.
The occupant information and feedback during the field studies are shown in Table 3b. 6 out of 33 studies provided relevant information. The information and feedback include occupant demographic information (e.g., living conditions), socio-economic status (e.g., highest qualifications), and their attitudes towards IAQ. The online scoring system, questionnaires, and interviews are the commonly adopted methods used to collect the relevant data.
The studies analyzed and included in this review are reported from different locations of the world. Out of the 33 studies, 9 (27.27%) field studies were carried out in the United States, while the rest of the studies are scattered across different countries or regions, including China, Australia, and the European Union. Possible reasons that explain why more studies have been done in the United States is that the UIAQS technology is more advanced and acknowledged in the United States. Figure 4 represents the distribution of included studies throughout the world.
Figure 5 illustrates the main IAQ sensor types reviewed in this study. Seven types of IAQ sensors can be categorized into three groups. Specifically, sensors that take up more than two-third of the total analyzed IAQ sensing systems are in Group A. The most prevalent type of sensors-PM2.5 sensors are the only sensor type that can be included in Group A, which account for 78.79% (26 out of 33 articles) of the total IAQ sensing systems. Group B includes the types of IAQ sensors that contribute to more than one-third and less than two-thirds of the analyzed IAQ sensing systems. In particular, sensors detecting TVOCs, CO2, and CO are in Group B, which hold 45.45%, 54.55%, and 45.45%, respectively. The last group—Group C—are those types of sensor that take up less than one-third of all the IAQ sensing systems, which include NO2 and O3 sensors.

6. Research Gaps and Suggestions

This section covers a detailed discussion of four major research limitations of current UIAQS technology deployment and proposes potential solutions for further studies.

6.1. Research Gaps of Current UIAQS Technologies

  • Sensing IAQ factors
It can be seen from Figure 5 that the distribution of UIAQ sensing factors is uneven. PM2.5 sensors take the leading position that more than 75% of the selected studies have PM2.5 sensors, while the sensing systems testing ozone concentration only account for 15.15% of the whole. None of the reviewed UIAQ sensing systems have a holistic measurement metric of odor, which is essential for IAQ-related factors that directly impact occupant perception of IAQ in the built environments. The lack of innovative IAQ sensor systems capturing some of the pertinent air pollutants, such as PM1, and sub-micron particles, may bring the issue of cross-sensitivity of sensors and the inability to monitor the holistic IAQ.
  • Lack of standardization
The second research gap is that the calibration of the UIAQ sensors has not been standardized, which decreases the accuracy of the measurement. The lack of standardization also indicates that the sensor performance is hard to compare with that of the reference sensors [76,81,82].
  • Applicability of UIAQS systems in extreme climates
From the geographic perspective, a vast number of studies are conducted in the temperate climate zone. The applicability of the UIAQS technologies to different regions has not been fully evaluated. Therefore, the implications of UIAQS systems drawn from the literature may not be applicable to buildings that are located in different climate zones. For instance, the humidity level in tropical areas is different from that of temperate zones, which has significant effects on the performance of UIAQS systems. Specifically, previous studies demonstrated a positive relationship between RH and sensor output (e.g., [88,89]).
  • Inadequate study duration
It has been found that the majority of field studies were conducted for relatively short durations, ranging from a couple of days to a few months. The evaluation of UIAQ sensors with adequate periods (i.e., over one year) is lacking in the literature. However, given the potential seasonal effects on sensor performance, research on the performance of UIAQS systems in field studies over a full year is definitely scarce.

6.2. Future Research

Based on the research limitations from the review, this study provides four suggestions for future directions of the development of UIAQS technologies. First of all, as suggested by prior studies (e.g., [56,61,66]), sensors detecting less pertinent IAQ pollutants and other IAQ-related factors, such as odor and humidity, are needed to be integrated into personalized UIAQS systems. As occupants’ perception of odors may vary from each other, no sensors have been deployed to clarify odors in a shared space. However, personalized UIAQS systems with the integration of occupant feedback of IAQ through questionnaires or interviews via interfaces have great potential to provide holistic IAQ measurements, including odors and other IAQ-related factors. In addition, with the emergence of COVID-19, it is crucial to explore the possibility of virus detection based on a long-range wireless sensor network that includes UIAQ sensors and a smart platform with analytics for airborne virus management [75]. Secondly, we suggest that researchers shall consider standardizing the UIAQS system calibration processes, as one of the major challenges identified is the lack of standardized and in-device calibration [57,68]. Thirdly, improving the diversification of the location of deployed UIAQ sensors is highly encouraged. The evaluation of UIAQ sensing performance can be undertaken in various locations across different regions of various climates. It is important to ensure that the performance of UIAQ sensors is consistent across different geographic locations. Last but not least, the evaluation of UIAQ sensors with adequate periods (i.e., over one year) is lacking in the literature [90]. Given the potential seasonal effects on sensor performance, the durations of IAQ testing studies are suggested to be extended for at least one year to account for seasonal effects [90]. The detailed development of integrating UIAQS systems and controls with human perception is elaborated in Section 7.

7. Integrating Air Quality Sensing and Controls with Human Perceptions

Relatively high airspeeds (generated by ventilation systems) and humidity impact both IAQ and thermal comfort at the same time. In most buildings nowadays, however, we have relatively low air speeds and an approximately fixed humidity, and the interrelationships become less prominent. In this case, thermal comfort is primarily impacted by air temperature [91], while IAQ is dominated by air pollutants and odor. To optimize IAQ, thermal comfort, and building energy simultaneously, we may need to adjust airspeeds, humidity rates and temperature accordingly to occupants’ instant needs, and that requires more sophisticated and comprehensive control systems, which require utilizing personalized sensors and actuators [92,93,94,95].
Following the future directions analyzed above, this section discusses how UIAQS technologies can holistically improve occupant health and thermal comfort spatiotemporally, as well as building energy efficiency through personalization of the IAQ. Personalization in this context refers to the development of IAQ sensing systems that can satisfy occupant requirements at the individual level. As illustrated in Figure 6, the PIAQS system requires the integration of occupant perception, ubiquitous IAQ sensors, and IAQ controls. Specifically, non-perceptible indoor air pollutants could be measured by physical sensors and occupant perception and feedback of IAQ can be captured via a user interface. The occupant feedback, in turn, may help adjust IAQ controls and ventilation.
To modify the micro-climate and provide a preferable indoor environment for occupants, IAQ controls and ventilation are essential in the PIAQS systems. Based on different types of UIAQ technologies, various solutions of IAQ controls can be adopted in personalized UIAQS systems [96]. For the stationary UIAQS systems, mechanical filtration [97], biofiltration [98], UV-C photolysis [99], UV-PCO [99], and air ionizers [100] are the most commonly used IAQ control methods. As for the mobile UIAQS technologies, except for biofiltration, other solutions applicable to the stationary UIAQ systems are acceptable. Specifically, mechanical filtration refers to the process of forcing air through fibrous media. Botanical biofiltration uses a mixture of plant species to remove targeted pollutants in the ambient air. UV-C Photolysis has frequently been used to inactivate microbes, such as SARS-CoV-2-virus, in recent years. UV-PCO is a process that seeks to remove TVOCs and NO from the air using oxidation, which can be incorporated into the ducting of HVAC systems or air purifiers [101]. Air ionizers make use of the process of ionization to clean the air, which can also be integrated into ventilation ducts of an HVAC system [102]. It is suggested that ion generators can remove small particles from the indoor air and improve occupant perception of IAQ. However, the application of ion generators in IAQ control is still controversial, as they are not effective at removing gases, odor, and large particles [103].
Ventilation is another essential component to improve IAQ in the built environment. One of the purposes of ventilation is to introduce fresh air into space by diluting indoor pollutant concentration levels [104]. Appropriate and adequate ventilation can help reduce the spread of viruses and microbiological contaminants at the individual level [105]. Personalized ventilation can adopt natural ventilation and mechanical ventilation. Natural ventilation uses wind and buoyancy to provide fresh air and to potentially increase air exchange rates [106,107]. However, due to the unpredictability of wind, natural ventilation is suggested to be combined with mechanical ventilation, which usually adopts fan-driven systems to provide clean air to occupants [108]. In extreme climates with very low or high humidity (e.g., deserts or tropical climates), natural ventilation may cause humidity-related challenges.

8. Conclusions

With the recent advancements in the IAQ sensing technologies, UIAQ systems are increasingly becoming instrumental in ensuring an individual’s health and comfort in the built environments. This study first critically reviewed the physical environmental factors that affect IAQ and the primary IAQ standards. It was discovered that the non-pollutant IAQ-related factors, such as odor, humidity, and microbes, are not properly addressed in the IAQ standards. However, these factors may have a diverse impact on occupant health, productivity, and their perception of indoor air quality. We then thoroughly reviewed the IAQ sensing standards, including US EPA, WELL v2, and RESET v2. It has been found that the guidelines for IAQ-sensing technologies are not standardized, and the requirements for different IAQ-related factors are varied. In addition, the sensing requirements on non-pollutant IAQ-related factors are exclusive in the guidelines for IAQ sensing technologies as well.
Accordingly, we systematically selected studies on ubiquitous sensing technologies deployed in the field studies from the literature. Of the relevant journal articles, 33 out of 229 distributed in 13 countries were selected for deeper analysis. We then extracted and analyzed several information items such as “study location”, “deployment”, “duration”, “measured factors”, “sensor modules”, “measurement range” “accuracy”, “sensor type”, and “occupant information and feedback”. We found that the majority of the UIAQ sensors could not satisfy the prominent IAQ sensing standards. Among the sensors, CO2 sensors have the highest percentage of 77.78% and 38.89% meeting the requirements of US EPA and RESET v2, respectively. Then, based on the UIAQS technologies reviewed in the literature, we identified four major research gaps in the applicability of current UIAQS technologies, including (1) sensing IAQ factors; (2) accuracy shortcoming; (3) applicability of UIAQS systems in extreme climates; and (4) inadequate study duration. Accordingly, four specific suggestions for future research directions are proposed to further improve UIAQS systems in an indoor environment, which include integrating personalization into UIAQS systems.
This study provides a critical reference for both researchers and practitioners studying IAQ sensing technologies in an indoor environment. The proposed “PIAQS” system provides a potential solution to provide personalized experience to the occupants. Researchers can utilize the concept of “personalization” in the context of the indoor environment and provide customized IAQ sensing and control strategies by collecting occupant perception and feedback. Policymakers are suggested to consider non-pollutant IAQ-related factors into the IAQ sensing standards to better guide the development of IAQ sensing technologies. Future research may focus on addressing the challenges of PIAQS technologies, thereby extending and enriching its application in practice.

Author Contributions

Conceptualization, Q.X., A.G., Z.V., Y.P. and H.C.G.; methodology, Q.X., A.G., E.M. and H.N.R.; validation, Q.X., A.G., E.M. and H.N.R.; formal analysis, Q.X., A.G. and H.C.G.; investigation, Q.X., A.G. and H.C.G.; resources, A.G., Z.V. and Y.P.; data curation, Q.X., A.G. and H.C.G.; writing—original draft preparation, Q.X., A.G. and H.C.G.; writing—review and editing, E.M. and H.N.R.; visualization, Q.X., A.G. and H.C.G.; supervision, A.G.; project administration, A.G., Z.V. and Y.P.; funding acquisition, A.G., Z.V. and Y.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Singapore Ministry of Education under Grant No. R-296-000-211-133 and Trane Technologies PLC under Grant No. A-0008468-00-00.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors wish to thank the financial support. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the Singapore Ministry of Education and Trane Technologies PLC.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Caron, A.; Redon, N.; Thevenet, F.; Hanoune, B.; Coddeville, P. Performances and limitations of electronic gas sensors to investigate an indoor air quality event. Build. Environ. 2016, 107, 19–28. [Google Scholar] [CrossRef]
  2. Joshi, S.M. The sick building syndrome. Indian J. Occup. Environ. Med. 2008, 12, 61–64. [Google Scholar] [CrossRef] [PubMed]
  3. Al Horr, Y.; Arif, M.; Kaushik, A.; Mazroei, A.; Katafygiotou, M.; Elsarrag, E. Occupant productivity and office indoor environment quality: A review of the literature. Build. Environ. 2016, 105, 369–389. [Google Scholar] [CrossRef] [Green Version]
  4. Khovalyg, D.; Kazanci, O.B.; Halvorsen, H.; Gundlach, I.; Bahnfleth, W.P.; Toftum, J.; Olesen, B.W. Critical review of standards for indoor thermal environment and air quality. Energy Build. 2020, 213, 109819. [Google Scholar] [CrossRef]
  5. Abdul-Wahab, S.A.; Chin Fah En, S.; Elkamel, A.; Ahmadi, L.; Yetilmezsoy, K. A review of standards and guidelines set by international bodies for the parameters of indoor air quality. Atmos. Pollut. Res. 2015, 6, 751–767. [Google Scholar] [CrossRef]
  6. Johnson, A.J. Cognitive facilitation following intentional odor exposure. Sensors 2011, 11, 5469–5488. [Google Scholar] [CrossRef] [Green Version]
  7. Wolkoff, P. Indoor air humidity, air quality, and health—An overview. Int. J. Hyg. Environ. Health 2018, 221, 376–390. [Google Scholar] [CrossRef]
  8. Wu, F.; Jacobs, D.; Mitchell, C.; Miller, D.; Karol, M.H. Improving indoor environmental quality for public health: Impediments and policy recommendations. Environ. Health Perspect. 2007, 115, 953–957. [Google Scholar] [CrossRef] [Green Version]
  9. Bhattacharya, A.; Ghahramani, A.; Mousavi, E. The effect of door opening on air-mixing in a positively pressurized room: Implications for operating room air management during the COVID outbreak. J. Build. Eng. 2021, 44, 102900. [Google Scholar] [CrossRef]
  10. Zhang, H.; Srinivasan, R. A systematic review of air quality sensors, guidelines, and measurement studies for indoor air quality management. Sustainability 2020, 12, 9045. [Google Scholar] [CrossRef]
  11. Coulby, G.; Clear, A.; Jones, O.; Godfrey, A. A scoping review of technological approaches to environmental monitoring. Int. J. Environ. Res. Public Health 2020, 17, 3995. [Google Scholar] [CrossRef] [PubMed]
  12. Saini, J.; Dutta, M.; Marques, G. A comprehensive review on indoor air quality monitoring systems for enhanced public health. Sustain. Environ. Res. 2020, 30, 6. [Google Scholar] [CrossRef] [Green Version]
  13. Chojer, H.; Branco, P.T.B.S.; Martins, F.G.; Alvim-Ferraz, M.C.M.; Sousa, S.I.V. Development of low-cost indoor air quality monitoring devices: Recent advancements. Sci. Total Environ. 2020, 727, 138385. [Google Scholar] [CrossRef] [PubMed]
  14. Morawska, L.; Thai, P.K.; Liu, X.; Asumadu-Sakyi, A.; Ayoko, G.; Bartonova, A.; Bedini, A.; Chai, F.; Christensen, B.; Dunbabin, M.; et al. Applications of low-cost sensing technologies for air quality monitoring and exposure assessment: How far have they gone? Environ. Int. 2018, 116, 286–299. [Google Scholar] [CrossRef]
  15. Vahlsing, C.; Smith, K.R. Global review of national ambient air quality standards for PM10 and SO2 (24 h). Air Qual. Atmos. Health 2012, 5, 393–399. [Google Scholar] [CrossRef] [Green Version]
  16. Davidson, C.I.; Phalen, R.F.; Solomon, P.A. Airborne Particulate Matter and Human Health: A Review. Aerosol Sci. Technol. 2005, 39, 737–749. [Google Scholar] [CrossRef]
  17. United States Environmental Protection Agency Particulate Matter (PM) Basics. Available online: https://www.epa.gov/pm-pollution/particulate-matter-pm-basics (accessed on 20 December 2021).
  18. Li, Z.; Wen, Q.; Zhang, R. Sources, health effects and control strategies of indoor fine particulate matter (PM2.5): A review. Sci. Total Environ. 2017, 586, 610–622. [Google Scholar] [CrossRef]
  19. Logue, J.M.; Price, P.N.; Sherman, M.H.; Singer, B.C. A method to estimate the chronic health impact of air pollutants in U.S. residences. Environ. Health Perspect. 2012, 120, 216–222. [Google Scholar] [CrossRef] [Green Version]
  20. Yang, B.-Y.; Guo, Y.; Morawska, L.; Bloom, M.S.; Markevych, I.; Heinrich, J.; Dharmage, S.C.; Knibbs, L.D.; Lin, S.; Yim, S.H.-L.; et al. Ambient PM1 air pollution and cardiovascular disease prevalence: Insights from the 33 Communities Chinese Health Study. Environ. Int. 2019, 123, 310–317. [Google Scholar] [CrossRef]
  21. World Health Organization (WHO). Air Quality Guidelines for Particulate Matter, Ozone, Nitrogen Dioxide and Sulphur Dioxide; Global Update 2005; World Health Organization: Geneva, Switzerland, 2006; Volume 38, p. E90038. [Google Scholar]
  22. World Health Organization; Regional Office for Europe. WHO Guidelines for Indoor Air Quality: Selected Pollutants; World Health Organization: Geneva, Switzerland; Regional Office for Europe: Copenhagen, Denmark, 2010; ISBN 978-92-890-0213-4. [Google Scholar]
  23. Chen, W.; Mendell, M.; Li, N.; Kumagai, K. Formaldehyde emissions from seams and cut edges of laminate flooring: Implications for emission testing protocols and exposure estimation. Build. Environ. 2018, 143, 652–660. [Google Scholar] [CrossRef]
  24. World Health Organization. Air Quality Guidelines for Europe; WHO Regional Office for Europe: Copenhagen, Denmark, 1987; ISBN 92-890-1114-9. [Google Scholar]
  25. Salthammer, T. Formaldehyde in the Ambient Atmosphere: From an Indoor Pollutant to an Outdoor Pollutant? Angew. Chem. Int. Ed. 2013, 52, 3320–3327. [Google Scholar] [CrossRef] [PubMed]
  26. Marzuki, P.F.; Abduh, M.; Driejana, R. Identification of source factors of carbon dioxide (CO2) emissions in concreting of reinforced concrete. Procedia Eng. 2015, 125, 692–698. [Google Scholar]
  27. Ghahramani, A.; Pantelic, J.; Vannucci, M.; Pistore, L.; Liu, S.; Gilligan, B.; Alyasin, S.; Arens, E.; Kampshire, K.; Sternberg, E. Personal CO2 bubble: Context-dependent variations and wearable sensors usability. J. Build. Eng. 2019, 22, 295–304. [Google Scholar] [CrossRef] [Green Version]
  28. Tran, V.V.; Park, D.; Lee, Y.-C. Indoor air pollution, related human diseases, and recent trends in the control and improvement of indoor air quality. Int. J. Environ. Res. Public. Health 2020, 17, 2927. [Google Scholar] [CrossRef] [Green Version]
  29. Bakó-Biró, Z.; Clements-Croome, D.J.; Kochhar, N.; Awbi, H.B.; Williams, M.J. Ventilation rates in schools and pupils’ performance. Build. Environ. 2012, 48, 215–223. [Google Scholar] [CrossRef]
  30. Jacobson, T.A.; Kler, J.S.; Hernke, M.T.; Braun, R.K.; Meyer, K.C.; Funk, W.E. Direct human health risks of increased atmospheric carbon dioxide. Nat. Sustain. 2019, 2, 691–701. [Google Scholar] [CrossRef]
  31. AirTest. A White Paper: CO2 Ventilation Control and Measurement of Outside Air; AirTest: Delta, BC, Canada, 2008; p. 5. [Google Scholar]
  32. Code of Practice for Indoor Air Quality for Air-Conditioned Buildings; Spring: Singapore, 2021.
  33. Ashrae, A. Standard 62.1-2019, Ventilation for Acceptable Indoor Air Quality; American Society of Heating, Refrigerating, and Air-Conditioning Engineers: Atlanta, GA, USA, 2019. [Google Scholar]
  34. Weaver, L.K. Carbon Monoxide Poisoning. N. Engl. J. Med. 2009, 360, 1217–1225. [Google Scholar] [CrossRef] [Green Version]
  35. National Research Council. Acute Exposure Guideline Levels for Selected Airborne Chemicals: Volume 8; National Academies Press: Washington, DC, USA, 2010; ISBN 0-309-14515-5. [Google Scholar]
  36. Environmental Protection Agency. Health Effects Associated with Nitrogen Dioxide. In Nitrogen Dioxide’s Impact on Indoor Air Quality; Environmental Protection Agency: Washington, DC, USA, 2021. [Google Scholar]
  37. Shuai, J.; Kim, S.; Ryu, H.; Park, J.; Lee, C.K.; Kim, G.-B.; Ultra, V.U.; Yang, W. Health risk assessment of volatile organic compounds exposure near Daegu dyeing industrial complex in South Korea. BMC Public Health 2018, 18, 528. [Google Scholar] [CrossRef] [Green Version]
  38. Zhai, S.; Jacob, D.J.; Wang, X.; Shen, L.; Li, K.; Zhang, Y.; Gui, K.; Zhao, T.; Liao, H. Fine particulate matter (PM2.5) trends in China, 2013–2018: Separating contributions from anthropogenic emissions and meteorology. Atmos. Chem. Phys. 2019, 19, 11031–11041. [Google Scholar] [CrossRef] [Green Version]
  39. Hoek, G.; Fischer, P.; van den Brandt, P.; Goldbohm, S.; Brunekreef, B. Estimation of long-term average exposure to outdoor air pollution for a cohort study on mortality. J. Expo. Sci. Environ. Epidemiol. 2001, 11, 459–469. [Google Scholar] [CrossRef] [Green Version]
  40. Hong, Y.-C.; Lee, J.-T.; Kim, H.; Kwon, H.-J. Air pollution: A new risk factor in ischemic stroke mortality. Stroke 2002, 33, 2165–2169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Turner, M.C.; Jerrett, M.; Pope, C.A., 3rd; Krewski, D.; Gapstur, S.M.; Diver, W.R.; Beckerman, B.S.; Marshall, J.D.; Su, J.; Crouse, D.L.; et al. Long-Term Ozone Exposure and Mortality in a Large Prospective Study. Am. J. Respir. Crit. Care Med. 2016, 193, 1134–1142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Silva, M.F.; Maas, S.; de Souza, H.A.; Gomes, A.P. Post-occupancy evaluation of residential buildings in Luxembourg with centralized and decentralized ventilation systems, focusing on indoor air quality (IAQ). Assessment by questionnaires and physical measurements. Energy Build. 2017, 148, 119–127. [Google Scholar] [CrossRef]
  43. Nagda, N.; Hodgson, M. Low relative humidity and aircraft cabin air quality. Indoor Air 2001, 11, 200–214. [Google Scholar] [CrossRef]
  44. Wolkoff, P.; Kjærgaard, S.K. The dichotomy of relative humidity on indoor air quality. Environ. Int. 2007, 33, 850–857. [Google Scholar] [CrossRef]
  45. Kim, M.; Chun, C.; Han, J. A Study on Bedroom Environment and Sleep Quality in Korea. Indoor Built Environ. 2010, 19, 123–128. [Google Scholar] [CrossRef]
  46. Okamoto-Mizuno, K.; Mizuno, K.; Michie, S.; Maeda, A.; Lizuka, S. Effects of Humid Heat Exposure on Human Sleep Stages and Body Temperature. Sleep 1999, 22, 767–773. [Google Scholar] [CrossRef] [Green Version]
  47. Sun, Y.; Zhang, Y.; Sundell, J.; Fan, Z.; Bao, L. Dampness in dorm rooms and its associations with allergy and airways infections among college students in China: A cross-sectional study. Indoor Air 2009, 19, 348. [Google Scholar] [CrossRef]
  48. Melikov, A.K.; Skwarczynski, M.A.; Kaczmarczyk, J.; Zabecky, J. Use of personalized ventilation for improving health, comfort, and performance at high room temperature and humidity. Indoor Air 2013, 23, 250–263. [Google Scholar] [CrossRef]
  49. Memarzadeh, F. Literature Review of the Effect of Temperature and Humidity on Viruses. ASHRAE Trans. 2012, 118, 1049–1060. [Google Scholar]
  50. Lax, S.; Smith, D.; Sangwan, N.; Handley, K.; Larsen, P.; Richardson, M.; Taylor, S.; Landon, E.; Alverdy, J.; Siegel, J. Colonization and succession of hospital-associated microbiota. Sci. Transl. Med. 2017, 9, eaah6500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Moelling, K.; Broecker, F. Air Microbiome and Pollution: Composition and Potential Effects on Human Health, Including SARS Coronavirus Infection. J. Environ. Public Health 2020, 2020, 1646943. [Google Scholar] [CrossRef] [PubMed]
  52. Bourdrel, T.; Annesi-Maesano, I.; Alahmad, B.; Maesano, C.N.; Bind, M.-A. The impact of outdoor air pollution on COVID-19: A review of evidence from in vitro, animal, and human studies. Eur. Respir. Rev. 2021, 30, 200242. [Google Scholar] [CrossRef] [PubMed]
  53. Williams, R.; Kilaru, V.; Snyder, E.; Kaufman, A.; Dye, T.; Rutter, A.; Russell, A.; Hafner, H. Air Sensor Guidebook; U.S. Environmental Protection Agency: Washington, DC, USA, 2014.
  54. WELL v2.0 Air Quality Monitoring and Awareness Part 1. 2018. Available online: https:/v2.wellcertified.com/v2.1/en/air/feature (accessed on 20 December 2021).
  55. Air Standards for Accredited Monitors v2.0. 2017. Available online: https://www.reset.build/standard/air (accessed on 20 December 2021).
  56. Curto, A.; Donaire-Gonzalez, D.; Barrera-Gómez, J.; Marshall, J.D.; Nieuwenhuijsen, M.J.; Wellenius, G.A.; Tonne, C. Performance of low-cost monitors to assess household air pollution. Environ. Res. 2018, 163, 53–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Abraham, S.; Li, X. A cost-effective wireless sensor network system for indoor air quality monitoring applications. Procedia Comput. Sci. 2014, 34, 165–171. [Google Scholar] [CrossRef] [Green Version]
  58. Baldelli, A. Evaluation of a low-cost multi-channel monitor for indoor air quality through a novel, low-cost, and reproducible platform. Meas. Sens. 2021, 17, 100059. [Google Scholar] [CrossRef]
  59. Barkjohn, K.K.; Bergin, M.H.; Norris, C.; Schauer, J.J.; Zhang, Y.; Black, M.; Hu, M.; Zhang, J. Using Low-cost sensors to Quantify the Effects of Air Filtration on Indoor and Personal Exposure Relevant PM2.5 Concentrations in Beijing, China. Aerosol Air Qual. Res. 2020, 20, 297–313. [Google Scholar] [CrossRef]
  60. Casey, J.G.; Ortega, J.; Coffey, E.; Hannigan, M. Low-cost measurement techniques to characterize the influence of home heating fuel on carbon monoxide in Navajo homes. Sci. Total Environ. 2018, 625, 608–618. [Google Scholar] [CrossRef]
  61. Coleman, J.; Teitelbaum, E.; Guo, H.; Read, J.; Meggers, F. Examining Architectural Air and Temperature with Novel Sensing Techniques. CISBAT 2017 Int. Conf. Build. Dist.-Energy Effic. Nano Urban Scale 2017, 122, 1135–1140. [Google Scholar] [CrossRef]
  62. Coulby, G.; Clear, A.K.; Jones, O.; Godfrey, A. Low-cost, multimodal environmental monitoring based on the Internet of Things. Build. Environ. 2021, 203, 108014. [Google Scholar] [CrossRef]
  63. Demanega, I.; Mujan, I.; Singer, B.C.; Anđelković, A.S.; Babich, F.; Licina, D. Performance assessment of low-cost environmental monitors and single sensors under variable indoor air quality and thermal conditions. Build. Environ. 2021, 187, 107415. [Google Scholar] [CrossRef]
  64. Gillooly, S.E.; Zhou, Y.; Vallarino, J.; Chu, M.T.; Michanowicz, D.R.; Levy, J.I.; Adamkiewicz, G. Development of an in-home, real-time air pollutant sensor platform and implications for community use. Environ. Pollut. 2019, 244, 440–450. [Google Scholar] [CrossRef] [PubMed]
  65. Kim, J.; Chu, C.; Shin, S. ISSAQ: An Integrated Sensing Systems for Real-Time Indoor Air Quality Monitoring. IEEE Sens. J. 2014, 14, 4230–4244. [Google Scholar] [CrossRef]
  66. Karami, M.; McMorrow, G.V.; Wang, L. Continuous monitoring of indoor environmental quality using an Arduino-based data acquisition system. J. Build. Eng. 2018, 19, 412–419. [Google Scholar] [CrossRef]
  67. Marinov, M.B.; Ganev, B.T.; Nikolov, D.N. Indoor Air Quality Assessment Using Low-cost Commercial Off-the-Shelf Sensors. In Proceedings of the 2021 6th International Symposium on Environment-Friendly Energies and Applications (EFEA), Sofia, Bulgaria, 24–26 March 2021; pp. 1–4. [Google Scholar]
  68. Manibusan, S.; Mainelis, G. Performance of Four Consumer-grade Air Pollution Measurement Devices in Different Residences. Aerosol Air Qual. Res. 2020, 20, 217–230. [Google Scholar] [CrossRef] [Green Version]
  69. Marques, G.; Pitarma, R. A Cost-Effective Air Quality Supervision Solution for Enhanced Living Environments through the Internet of Things. Electronics 2019, 8, 170. [Google Scholar] [CrossRef] [Green Version]
  70. Mazaheri, M.; Clifford, S.; Yeganeh, B.; Viana, M.; Rizza, V.; Flament, R.; Buonanno, G.; Morawska, L. Investigations into factors affecting personal exposure to particles in urban microenvironments using low-cost sensors. Environ. Int. 2018, 120, 496–504. [Google Scholar] [CrossRef] [PubMed]
  71. Moreno-Rangel, A.; Sharpe, T.; Musau, F.; McGill, G. Field evaluation of a low-cost indoor air quality monitor to quantify exposure to pollutants in residential environments. J. Sens. Sens. Syst. 2018, 7, 373–388. [Google Scholar] [CrossRef] [Green Version]
  72. Palmisani, J.; Di Gilio, A.; Viana, M.; de Gennaro, G.; Ferro, A. Indoor air quality evaluation in oncology units at two European hospitals: Low-cost sensors for TVOCs, PM2.5 and CO2 real-time monitoring. Build. Environ. 2021, 205, 108237. [Google Scholar] [CrossRef]
  73. Parkinson, T.; Parkinson, A.; de Dear, R. Continuous IEQ monitoring system: Performance specifications and thermal comfort classification. Build. Environ. 2019, 149, 241–252. [Google Scholar] [CrossRef] [Green Version]
  74. Patel, S.; Li, J.; Pandey, A.; Pervez, S.; Chakrabarty, R.K.; Biswas, P. Spatio-temporal measurement of indoor particulate matter concentrations using a wireless network of low-cost sensors in households using solid fuels. Environ. Res. 2017, 152, 59–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Peladarinos, N.; Cheimaras, V.; Piromalis, D.; Arvanitis, K.G.; Papageorgas, P.; Monios, N.; Dogas, I.; Stojmenovic, M.; Tsaramirsis, G. Early Warning Systems for COVID-19 Infections Based on Low-Cost Indoor Air-Quality Sensors and LPWANs. Sensors 2021, 21, 6183. [Google Scholar] [CrossRef] [PubMed]
  76. Pokhrel, A.K.; Bates, M.N.; Acharya, J.; Valentiner-Branth, P.; Chandyo, R.K.; Shrestha, P.S.; Raut, A.K.; Smith, K.R. PM2.5 in household kitchens of Bhaktapur, Nepal, using four different cooking fuels. Atmos. Environ. 2015, 113, 159–168. [Google Scholar] [CrossRef]
  77. Qabbal, L.; Younsi, Z.; Naji, H. An indoor air quality and thermal comfort appraisal in a retrofitted university building via low-cost smart sensor. Indoor Built Environ. 2021, 31, 586–606. [Google Scholar] [CrossRef]
  78. Semple, S.; Ibrahim, A.E.; Apsley, A.; Steiner, M.; Turner, S. Using a new, low-cost air quality sensor to quantify second-hand smoke (SHS) levels in homes. Tob. Control 2015, 24, 153–158. [Google Scholar] [CrossRef]
  79. Spinelle, L.; Gerboles, M.; Villani, M.G.; Aleixandre, M.; Bonavitacola, F. Field calibration of a cluster of low-cost available sensors for air quality monitoring. Part A: Ozone and nitrogen dioxide. Sens. Actuators B Chem. 2015, 215, 249–257. [Google Scholar] [CrossRef]
  80. Spinelle, L.; Gerboles, M.; Villani, M.G.; Aleixandre, M.; Bonavitacola, F. Field calibration of a cluster of low-cost commercially available sensors for air quality monitoring. Part B: NO, CO and CO2. Sens. Actuators B Chem. 2017, 238, 706–715. [Google Scholar] [CrossRef]
  81. Steinle, S.; Reis, S.; Sabel, C.E.; Semple, S.; Twigg, M.M.; Braban, C.F.; Leeson, S.R.; Heal, M.R.; Harrison, D.; Lin, C.; et al. Personal exposure monitoring of PM2.5 in indoor and outdoor microenvironments. Sci. Total Environ. 2015, 508, 383–394. [Google Scholar] [CrossRef] [Green Version]
  82. Tiele, A.; Esfahani, S.; Covington, J. Design and Development of a Low-Cost, Portable Monitoring Device for Indoor Environment Quality. J. Sens. 2018, 2018, 5353816. [Google Scholar] [CrossRef]
  83. Trilles, S.; Vicente, A.B.; Juan, P.; Ramos, F.; Meseguer, S.; Serra, L. Reliability Validation of a Low-Cost Particulate Matter IoT Sensor in Indoor and Outdoor Environments Using a Reference Sampler. Sustainability 2019, 11, 7220. [Google Scholar] [CrossRef] [Green Version]
  84. Wang, Z.; Delp, W.W.; Singer, B.C. Performance of low-cost indoor air quality monitors for PM2.5 and PM10 from residential sources. Build. Environ. 2020, 171, 106654. [Google Scholar] [CrossRef]
  85. Zamora, M.L.; Rice, J.; Koehler, K. One year evaluation of three low-cost PM2.5 monitors. Atmos. Environ. 2020, 235, 117615. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, H.; Srinivasan, R.; Ganesan, V. Low Cost, Multi-Pollutant Sensing System Using Raspberry Pi for Indoor Air Quality Monitoring. Sustainability 2021, 13, 370. [Google Scholar] [CrossRef]
  87. Zhao, Y.; Zhao, B. Reducing human exposure to PM2.5 generated while cooking typical Chinese cuisine. Build. Environ. 2020, 168, 106522. [Google Scholar] [CrossRef]
  88. Tryner, J.; L’Orange, C.; Mehaffy, J.; Miller-Lionberg, D.; Hofstetter, J.C.; Wilson, A.; Volckens, J. Laboratory evaluation of low-cost PurpleAir PM monitors and in-field correction using co-located portable filter samplers. Atmos. Environ. 2020, 220, 117067. [Google Scholar] [CrossRef]
  89. Zou, Y.; Young, M.; Chen, J.; Liu, J.; May, A.; Clark, J.D. Examining the functional range of commercially available low-cost airborne particle sensors and consequences for monitoring of indoor air quality in residences. Indoor Air 2020, 30, 213–234. [Google Scholar] [CrossRef] [PubMed]
  90. Nguyen, J.L.; Schwartz, J.; Dockery, D.W. The relationship between indoor and outdoor temperature, apparent temperature, relative humidity, and absolute humidity. Indoor Air 2014, 24, 103–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Arens, E.; Ghahramani, A.; Przybyla, R.; Andersen, M.; Min, S.; Peffer, T.; Raftery, P.; Zhu, M.; Luu, V.; Zhang, H. Measuring 3D indoor air velocity via an inexpensive low-power ultrasonic anemometer. Energy Build. 2020, 211, 109805. [Google Scholar] [CrossRef] [Green Version]
  92. Bhattacharya, A.; Pantelic, J.; Ghahramani, A.; Mousavi, E.S. Three-dimensional analysis of the effect of human movement on indoor airflow patterns. Indoor Air 2021, 31, 587–601. [Google Scholar] [CrossRef]
  93. Ghahramani, A.; Zhu, M.; Przybyla, R.; Andersen, M.; Min, S.; Zhang, H.; Peffer, T.; Arens, E. An Inexpensive Low-Power Ultrasonic 3-Dimensional Air Velocity Sensor. In Proceedings of the 2019 IEEE Sensors, Montreal, QC, Canada, 27–30 October 2019; pp. 1–4. [Google Scholar]
  94. Ghahramani, A.; Zhang, K.; Dutta, K.; Yang, Z.; Becerik-Gerber, B. Energy savings from temperature setpoints and deadband: Quantifying the influence of building and system properties on savings. Appl. Energy 2016, 165, 930–942. [Google Scholar] [CrossRef] [Green Version]
  95. Ghahramani, A.; Galicia, P.; Lehrer, D.; Varghese, Z.; Wang, Z.; Pandit, Y. Artificial intelligence for efficient thermal comfort systems: Requirements, current applications and future directions. Front. Built Environ. 2020, 6, 49. [Google Scholar] [CrossRef]
  96. Ghahramani, A.; Tang, C.; Becerik-Gerber, B. An online learning approach for quantifying personalized thermal comfort via adaptive stochastic modeling. Build. Environ. 2015, 92, 86–96. [Google Scholar] [CrossRef] [Green Version]
  97. Kim, J.-H.; Kim, H.-G.; Yeo, M.-S. Ventilation and Filtration Control Strategy Considering PM2.5, IAQ, and System Energy. Atmosphere 2020, 11, 1140. [Google Scholar] [CrossRef]
  98. Rodgers, K.; Handy, R.; Hutzel, W. Indoor air quality (IAQ) improvements using biofiltration in a highly efficient residential home. J. Green Build. 2013, 8, 22–27. [Google Scholar] [CrossRef]
  99. Zhong, L.; Haghighat, F.; Lee, C.-S.; Lakdawala, N. Performance of ultraviolet photocatalytic oxidation for indoor air applications: Systematic experimental evaluation. J. Hazard. Mater. 2013, 261, 130–138. [Google Scholar] [CrossRef]
  100. Seraj, M. Managing and Monitoring IndoorAir Quality Using bi-polarAir Ionizer. Invertis J. Sci. Technol. 2019, 12, 95–109. [Google Scholar]
  101. Zhong, L.; Haghighat, F.; Lee, C.-S. Ultraviolet photocatalytic oxidation for indoor environment applications: Experimental validation of the model. Build. Environ. 2013, 62, 155–166. [Google Scholar] [CrossRef]
  102. Schurk, D.N. A Bipolar Ionization Primer for HVAC Professionals. ASHRAE J. 2021, 63, 40–46. [Google Scholar]
  103. Environmental Protection Agency. What are ionizers and other ozone generating air cleaners? In Indoor Air Pollution: An Introduction for Health Professionals; Environmental Protection Agency: Washington, DC, USA, 2021. [Google Scholar]
  104. Xue, Q.; Wang, Z.; Liu, J.; Dong, J. Indoor PM2. 5 concentrations during winter in a severe cold region of China: A comparison of passive and conventional residential buildings. Build. Environ. 2020, 180, 106857. [Google Scholar] [CrossRef]
  105. Bolashikov, Z.D.; Melikov, A.K. Methods for air cleaning and protection of building occupants from airborne pathogens. Build. Environ. 2009, 44, 1378–1385. [Google Scholar] [CrossRef]
  106. Ghahramani, A.; Zhu, M.; Przybyla, R.J.; Andersen, M.P.; Galicia, P.J.; Peffer, T.E.; Zhang, H.; Arens, E. Measuring Air Speed With a Low-Power MEMS Ultrasonic Anemometer via Adaptive Phase Tracking. IEEE Sens. J. 2019, 19, 8136–8145. [Google Scholar] [CrossRef] [Green Version]
  107. Li, Y.; Delsante, A. Natural ventilation induced by combined wind and thermal forces. Build. Environ. 2001, 36, 59–71. [Google Scholar] [CrossRef]
  108. Chartier, Y.; Pessoa-Silva, C.L. Natural Ventilation for Infection Control in Health-Care Settings; WHO: Geneva, Switzerland, 2009. [Google Scholar]
Figure 1. Comparison of the accuracy requirements of three standards on IAQ sensing technologies. Note: As the accuracy requirements are not available in US EPA, minimum output resolution was adopted instead.
Figure 1. Comparison of the accuracy requirements of three standards on IAQ sensing technologies. Note: As the accuracy requirements are not available in US EPA, minimum output resolution was adopted instead.
Sensors 22 03444 g001
Figure 2. Review procedures.
Figure 2. Review procedures.
Sensors 22 03444 g002
Figure 3. Sensor modules of various IAQ pollutants and sensing requirements. Note: CO2 and CO2, PM2.5 and PM 2.5, PM10 and PM10, NO2 and NO2, and O3 and O3 are used interchangeably.
Figure 3. Sensor modules of various IAQ pollutants and sensing requirements. Note: CO2 and CO2, PM2.5 and PM 2.5, PM10 and PM10, NO2 and NO2, and O3 and O3 are used interchangeably.
Sensors 22 03444 g003
Figure 4. Geographic distribution of the reviewed studies in UIAQS technologies.
Figure 4. Geographic distribution of the reviewed studies in UIAQS technologies.
Sensors 22 03444 g004
Figure 5. Distribution of the tested IAQ factors. Note: CO2 and CO2, PM2.5 and PM2.5, PM10 and PM10, and NO2 and NO2 are used interchangeably.
Figure 5. Distribution of the tested IAQ factors. Note: CO2 and CO2, PM2.5 and PM2.5, PM10 and PM10, and NO2 and NO2 are used interchangeably.
Sensors 22 03444 g005
Figure 6. Personalized Indoor Air Quality Sensing (PIAQ) System.
Figure 6. Personalized Indoor Air Quality Sensing (PIAQ) System.
Sensors 22 03444 g006
Table 1. The potential health effects and recommended levels of indoor air pollutants.
Table 1. The potential health effects and recommended levels of indoor air pollutants.
Air PollutantPotential Health EffectsIAQ Code of Practice
ASHRAESingapore Standard SS 554WHO 2006 Guidelines
Carbon DioxideHeadache/Fatigue/Nausea/Dizziness1000 ppm700 ppm above outdoorN/A
Carbon MonoxideFatigue/Impaired vision/Reduced brain function/Nausea/Headaches/Dizziness/Flu-like symptoms/Fatal9 ppm9 ppm6 ppm
FormaldehydeAsthma/microvascular leakage/cancer0.1 ppm (office)0.1 ppm0.1 ppm
0.04 ppm (home)
Nitrogen DioxideEye, nose, throat irritation/Acute or chronic bronchitis/Respiratory infectionsN/AN/A200 μg/m3
OzoneRespiratory illness, such as cardiovascular mortalityN/A50 ppb60 ppb
TVOCsEye, nose and throat irritation/Nausea/Headaches, loss of coordination/Damage to liver, kidney, and central nervous system/Skin irritationN/A3000 ppbN/A
Particulate MatterEye, nose, and throat irritation/Aggravation of respiratory tract related ailments50 μg/m3 (PM10PM10)50 μg/m3 (PM10PM10)50 μg/m3 (PM10PM10)
25 μg/m3 (PM2.5PM2.5)
Table 2. Categorization of the selected UIAQS technology-related studies.
Table 2. Categorization of the selected UIAQS technology-related studies.
CategoryDescription
Study locationStudy location revealed the country, the location where the study was conducted (e.g., kitchen, living room, etc.). The rationale behind identifying the country is to grasp a better sense of the country’s climate.
DeploymentThe deployment column recorded the specific locations where the sensors were installed (e.g., on the desk, near fans, etc.). This information helps identify the various tested factors and the sensors’ suitability for different scenarios.
DurationDuration revealed the time length of the studies, such as seven days or one year.
Measured factorsThe measured factors recorded the type of indoor environmental factors that affect IAQ.
Sensor ModulesThe sensor modules column kept track of the sensor brand and model name.
Measurement rangeThe measurement range documented the measuring range of each environmental factor
AccuracyThe accuracy information was extracted from the specifications of the sensors, which helps in evaluating the suitability of the sensors when deployed in various settings.
Sensor typeSensor type indicated whether this UIAQ sensor is a stationary or a mobile sensor.
Occupant information and feedbackOccupant information and feedback took a record of the contextual information and feedback from the occupants via online feedback, questionnaires, and interviews.
Table 3. (a). Air quality sensing systems characteristics, deployment, and study information. (b). Summary of occupant information and feedback collected in the literature.
Table 3. (a). Air quality sensing systems characteristics, deployment, and study information. (b). Summary of occupant information and feedback collected in the literature.
(a)
StudyStudy
Location
DeploymentDurationMeasured
Factors
Sensor ModulesMeasurement
Range
AccuracySensor Type
[1]Laboratory
France
Sensor 1:
in front of green wall, 0.7 m above floor
Sensor 2 and 3: Near fans
Sensor 4:
in a corner of room
-COFigaro TGS244230–1000 ppm-S
CO2Figaro TGS4161350–10,000 ppm±20%
TVOCsFigaro TGS 26021–30 ppm-
O3E2v MICS 261010–10,000 ppb
NO2E2v MICS 27140.05–5 ppm
[56]Residential
Spain and India
Spain:
1 m from an indoor fireplace and 0.6 m above the ground
India:
main living area at least 1 m above the ground
Spain:
5 days
India:
7 days
COEL-USB-CO3–1000 ppm±7 ppmS
PM 2.5 HAPEx Nano 8 150   mg / m 3 -
[57]Laboratory USA--CO2MG811350–10,000 ppm-S
TVOCsTGS26021–30 ppm
COMQ720–2000 ppm
O3MQ13110–1000 ppb
[58]Residential
Canada
1 m above a drawer at the center of a room7 days PM 2.5 Shinyei Kaisha PPD42-60 0 200   μ g / m 3 ±20 μg/m3 or
10%
S
CO2ELT Sensor T-110-3V400 to 10,000 ppm±50 ppm or 3%
COFigaro Engineering TGS53420–1000 ppm±10 ppm
TVOCsCambridge CMOS CC881B0–1000 ppb±10 ppb or 5%
O3SGX Sensortech MICS-27140–1000 ppb±10 ppb or 5%
NO2SGX Sensortech MICS-27140–1000 ppb±10 ppb or
5%
[59]Residential
China
1 m above the ground4 days PM 2.5 PMS 3003 0 500   μ g / m 3 100 500   μ g / m 3 :
± 10 %   0 100   μ g / m 3 :
± 10   μ g / m 3
S
[60]Residential
USA
1 to 5 m from stove or furnace2 monthsCOAlphasense COB40–45 ppm±10 ppmS
[61]Laboratory
USA
--CO2Telaire T67130–5000 ppm±30 ppm
+3%
S
TVOCsMiCS-55241–1000 ppm-
COMiCS-55241–1000 ppm
PM 2.5 Itead DSM501A-
PM 10 Itead DSM501A-
FormaldehydeWSP21101–50 ppm
NO2MiCS-27140.05–10 ppm
[62]Office
UK
On a shelf (1.65 m high) above a computer desk1 monthTVOCsCCS-8110–1187 ppb-S
CO2MH-Z19B0–2000 ppm
PM 2.5 PMSA003i 0 500   μ g / m 3
HumidityBME2800–100%
[63]Laboratory
Switzerland
On the table at 0.75 m above the ground1 hCO2Sensirion SCD400–40000 ppm±50 ppm
±5%
S
CO2meter K300–10,000 ppm±30 ppm
±3%
PM 2.5   and   PM 10 NovaFitness SDS018 0 1000   μ g / m 3 max   of   ± 15 %
& ± 10 μg/m3
[64]Residential
USA
On a table in the main living space7 daysCOAlphasense COB40–1000 ppm±10 ppmS
CO2Netatmo0–5000 ppm±50 ppm,
±5%
NOAlphasense NOB40–20 ppm±10 ppb
NO2Alphasense NO2B43F0–20 ppm±50 ppb
PM 2.5 Alphasense OPC-N20.38–17 μm-
[65]Residential, campus and church
USA
Living room, Classroom, and Church-TVOCsTGS26021–30 ppm-S
NO2GSNT110–200 ppm±10%
COTGS50420–10,000 ppm-
SO2SO2-AF400–700 ppm
PM 2.5 GP2Y1010AUF-
PM 10 GP2Y1010AUF
[66]Laboratory
USA
-10 daysCO2CO2Meter K-300–10,000 ppm±30 ppm
± 3%
S
PM 2.5 GP2Y1010AU0F 0 0.5 mg / m 3 -
TVOCsCO2Meter IAQ-2000350–2000 ppm
[67]-Test chamber-CO2SenseAir K300–10,000 ppm±30 ppm
+3%
S
TVOCsAMS CCS8110–1200 ppb-
SGX Sensortech MICS-VZ89TE0–2000
Bosch BME680-
Bosch BME688
[68]Residential
USA
Near the edge of a stable surface in the center of the living room2 months PM 2.5 AQE2--S
BlueAir Aware 1 500   ug / m 3
Foobot 0 1300   µ g / m 3 ±4 µg or
±20%
Speck (DSM501A)--
[69]Laboratory
Portugal
--COMICS-68141–1000 ppm-S
NO2MICS-68150.05–10 ppm
C2H5OHMICS-681610–500 ppm
H2MICS-68171–1000 ppm
NH3MICS-68181–500 ppm
CH4MICS-6819>1000 ppm
C3H8MICS-6820>1000 ppm
C4H10MICS-6821>1000 ppm
[70]Residential and commercial
Australia
Personal monitor carried by people1 week PM 2.5 Shinyei PPD60PV-T2--M
[71]Residential
UK
0.9 m above a drawer4 days PM 2.5 SHARP GP2Y1010AU0F 0 1300   µ g / m 3 ± 4   µ g / m 3 S
TVOCsAMS iAQ-CORE-C125–1000 ppb±1.0 ppm
CO2AMS iAQ-CORE-C400–600 ppm±1.0 ppm
[72]Hospital
Italy and Spain
2.5 m and 1 m above the groundMultiple months PM 2.5 Syhitech DSM501A--S
Model 3321 Aerodynamic Particle Sizer
SPK202
SPK201
TVOCsCorvus0–50 ppm±5 ppb
CO2ZyAura0–3000 ppm±75 ppm or
±5%
[73]Commercial
Australia
In Office buildings-COTSI Q-Trak 7575500–2000 ppm±3% or
±50 ppm
S
CO2Fieldpiece SCM40–15 ppm±5% or
±1 ppm
PM 2.5 TSI DustTrak II 8532 0 0.1   μ g / m 3 -
PM 10 TSI DustTrak II 8532 0 0.1   μ g / m 3
FormaldehydeHalTech HFX2050–500 ppb
[74]Residential
India
Integrated in the kitchen (areas around the cookstove) and roomMultiple days PM 2.5 GP2Y1010AU0F--S
[75]Hospital1.2 m above the ground at the center of each room24 hTVOCsSensirion SVM300–60,000 ppb1.3%S
Renesas ZMOD44100–1,000,000 ppb±25%
CO2Sensirion SVM30400–60,000 ppm1.3%
Renesas ZMOD4410400–5000 ppm±25%
PM 2.5 Sensirion SPS30 1 1000   µ g / m 3 0 1000   µ g / m 3 :
± 10   µ g / m 3
PM 10 Sensirion SPS30 1 1000   µ g / m 3 0 100   µ g / m 3 :
± 25   µ g / m 3
100 1000   µ g / m 3 :
± 25%
[76]Residential
Nepal
1.5 m height and about 1 m from the edge of the main stove and ≥1 m from any doors or other openings in the walls13 months PM 2.5 UCB-PATS 25 25 , 000   µ g / m 3 -S
[77]Laboratory
France
1.4 m above the ground on the wall behind occupants8 monthsCO2NDIR0–5000 ppm50 ppmS
IQ 6100–10,000 ppm50 ppm
C6H6 andCOMQ13510–1000 ppm±5%
FormaldehydeMS11000–1000 ppm±3%
PM 2.5 GP2Y1010AU0F 0 500   mg / m 3 0.1   mg / m 3
[78]Residential
Scotland
1 m above the ground in living area, away from possible PM sources2 months PM 2.5 Dylos DC1700--S
[79]Laboratory
Italy
On an evaluation board5 monthsCOTGS-50420–10,000 ppm-S
MICS-4514-CO0–1000 ppm
CO2Gascard NG0–2000/3000/5000 ppm±2% of range ±<0.015% of range per mbar
S-1000–2000/3000/5000/10,000 ppm±30 ppm
±5%
NO2NO2B40–20 ppm-
NO2_3E500.3–50 ppm
MICS-27100.05–5 ppm
MICS-4514-NO20.05–10 ppm
CairClip NO20–250 ppb
O3O3B40–5 ppm
O3_3E1F0.1–1 ppm
NONO_3E1000–100/200 ppm45 nA/ ppm
± 15 nA/ ppm
[80]Laboratory
Italy
--NOCitytech NO_3E1000–1000 ppm45 ± 15
nA/ppm
S
COFigaro TGS-50420–10,000 ppm1.2–2.4
nA/ppm
e2V MICS-45140–1000 ppm−0.0051(Rs/R0)
/ppm
CO2Edinburgh Gascard0–1000 ppm1 V/100 ppm
ELT S-100H0–5000 ppm1 V/1000 ppm
[81]Residential and commercial
Scotland
Integrated in a backpack worn by subjects2 months PM 2.5 Dylos 1700--S
[82]Laboratory
UK
On a 0.75–1.2 m height desk in front of occupant’s work areas-HumiditySHT310–100%±2% RHS
PM 2.5 / PM 10 HPMA115S0 0 1000   μ g / m 3 ±15%
TVOC1CCS8110–1200 ppb-
TVOC2iAQ-CoreC125–600 ppb
TVOC3MiCS-VZ-89TE0–1000 ppb
CO2T67130–5000 ppm±25 ppm
COLLC110-1020–1000 ppm±2 ppm
[83]Office
Spain
Inside an office with no human presence- PM 2.5 HM-3301 1 1000   μ g / m 3 -S
PM 10 HM-3301 1 1000   μ g / m 3
[84]LaboratoryOn or adjacent to a wire shelving unit in the central area, several meters from source activities- PM 2.5   and   PM 10 Air Quality Egg 2018 (AQE) 0 500   μ g / m 3 0 100   μ g / m 3
:   ± 10   μ g / m3
100–500 μg/m3:
±10%
S
IQAir AirVisual Pro (AVP) 0 1798   μ g / m 3 -
Awair 2nd Edition (AW2)- 0 100   μ g / m 3 :
± 15   μ g / m 3   100 1000   μ g / m 3 :
±15%
Kaiterra Laser Egg 2 (LE2) 0 500   μ g / m 3 0 100   μ g / m 3 :
± 10   μ g / m 3
100–500 μg/m3:
±10%
PurpleAir Indoor (PAI) 0 500   μ g / m 3 0 100   μ g / m 3 :
± 10   μ g / m 3
100 500   μ g / m 3 :
±10%
Ikair (IKA) 0 1000   μ g / m 3 0 100   μ g / m 3 :
± 10   μ g / m 3
100 1000   μ g / m 3 :
±10%
[85]Residential
USA
On a table in the living room, 3 m from the main entrance and 10 m from the kitchen12 months PM 2.5 AirVisual Pro-±8%S
[86]Laboratory and residential
USA
1.1 m above the ground, 0.5 m away from a wall, and at least a 1.5 m away from any corner7 days PM 2.5   and   PM 10 Nova Fitness SDS011 0.0 999.9   µ g   / m 3 15 %   or   ± 10   µ g / m 3 S
NO2SPEC Sensors DGS-NO2 968-0430–10 ppm±15%
SO2SPEC Sensors DGS-SO2 968-0380–20 ppm±15%
CO2CO2Meter K-300–5000 ppm±30%
COSPEC Sensors DGS-CO 968-0340–1000 ppm±15%
O3SPEC Sensors DGS-O3 968-0420–5 ppm±15%
TVOCsOhmetech.io uThing:VOC™0–500 IAQ index±15%
[87]Residential
China
Sensor 1:
1.2 m above the ground in adjacent room
Sensor 2:
in a shoulder brace 0.15 m away from cook’s nose
1 month and 5 days PM 2.5 AM510--S
(b)
StudyOccupant Information and Feedback
[59]A brief survey about each home, its residents, and general behavior patterns, such as home parameters, cooking days, other potential indoor PM sources.
[70]Questionnaires on participants’ age and gender, socio-economic status (highest qualification) and characteristics of the indoor and outdoor environments during weekdays and weekend, such as age for their residence, ventilation conditions at home and work, type of cooktop at home and commuting preferences.
[76]Questionnaires about three groups of variables:
(1) Variables for the day of monitoring, such as type of primary and secondary stoves used in the house, unusual stove use pattern during the HAP monitoring period, weather conditions during the monitoring period, ventilation in the kitchen (e.g., open doors and windows) and other smoke exposure sources, such as number of smokers in the house, use of incense or mosquito coils.
(2) Fixed variables, such as kitchen size or the presence of roads within 100 m.
(3) Variables describing usual practices, such as types of non-electric lamp used when power is unavailable, and type of space heating used during the winter.
Interviews about participants’ caregivers or parents about their occupations and household characteristics, such as construction materials.
[78]Questionnaires on occupant attitudes towards smoking and interest in having a device placed at home to measure air quality.
[81]Interviews and online questionnaires on personal data, such as individual’s living conditions, the household size and accommodation details, building and neighborhood characteristics and other contextual factors.
[82]Feedback through an online IEQ scoring system.
Note: (1) Unreported: “-”. (2) Sensor Type: S (Stationary); M (Mobile).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xu, Q.; Goh, H.C.; Mousavi, E.; Nabizadeh Rafsanjani, H.; Varghese, Z.; Pandit, Y.; Ghahramani, A. Towards Personalization of Indoor Air Quality: Review of Sensing Requirements and Field Deployments. Sensors 2022, 22, 3444. https://doi.org/10.3390/s22093444

AMA Style

Xu Q, Goh HC, Mousavi E, Nabizadeh Rafsanjani H, Varghese Z, Pandit Y, Ghahramani A. Towards Personalization of Indoor Air Quality: Review of Sensing Requirements and Field Deployments. Sensors. 2022; 22(9):3444. https://doi.org/10.3390/s22093444

Chicago/Turabian Style

Xu, Qian, Hui Ci Goh, Ehsan Mousavi, Hamed Nabizadeh Rafsanjani, Zubin Varghese, Yogesh Pandit, and Ali Ghahramani. 2022. "Towards Personalization of Indoor Air Quality: Review of Sensing Requirements and Field Deployments" Sensors 22, no. 9: 3444. https://doi.org/10.3390/s22093444

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