Skip to main content

Effect of forceful suction and air disinfection machines on aerosol removal

A Correction to this article was published on 19 December 2023

This article has been updated

Abstract

Backgrounds

Dental procedures involving drilling and grinding can produce a significant amount of suspended aerosol particles (PM) and bioaerosols. This study aims to analyze the size and concentration of aerosol particles generated during drilling and to investigate the effectiveness of two air exchange systems, namely forceful suction (FS) and air disinfection machines (DM), in removing PM.

Methods

For this study, 100 extracted permanent teeth were collected and divided into three groups: without suction (n = 50), suction with forceful suction (n = 25), and suction with air disinfection machines (n = 25). The removal rate of suspended aerosol particles was analyzed using particle counters and air data multimeter.

Results

When drilling and grinding were performed without vacuum, 0.75% of the aerosol particles generated were PM2.5-10, 78.25% of total suspended aerosol particles (TSP) were PM2.5, and 98.68% of TSP were PM1. The nanoanalyzer measurements revealed that the aerodynamic diameter of most aerosol particles was below 60 nm, with an average particle diameter of 52.61 nm and an average concentration of 2.6*1011 ultrafine aerosol particles. The air change per hour (ACH) was significantly lower in the air disinfection machines group compared to the forceful suction group. Additionally, the number of aerosol particles and mass concentration was significantly lower in the air disinfection machines group compared to the forceful suction group in terms of PM2.5 levels. However, the forceful suction group also reduced the mass concentration in PM10 level than the air disinfection machines group.

Conclusion

In conclusion, the air exchange system can reduce the aerosol particles generated during drilling and grinding. Comparing the two air exchange systems, it was found that the air disinfection machines group reduces the number of aerosol particles and mass concentration in PM2.5 levels, while the forceful suction group reduces the mass concentration in PM10 level.

Peer Review reports

Introduction

Dental caries and periodontal disease are prevalent conditions that pose a significant threat to oral health [1, 2]. Common therapeutic interventions for these diseases encompass restorative dentistry, extractions, scaling, and endodontic treatments [3, 4]. Within the context of dental procedures, aerosols primarily emanate from oral fluids produced by patients, consequently leading to the dissemination of microorganisms and compromising the environmental quality of dental offices [5]. High-speed dental equipment utilization or rinsing during scaling primarily engenders the generation of these aerosols [6, 7], wherein various bacteria, such as Streptococcus spp., have been found in dental offices [8]. Importantly, even upon completion of treatment, noteworthy concentrations of airborne bacteria persist within the dental office vicinity, with potential for dissemination to other treatment areas [9]. Consequently, this scenario poses potential health risks for both patients and dental practitioners [10].

Furthermore, as a supplement to standard protective measures like wearing masks and gloves during dental treatments, the utilization of appropriate preoperative mouth rinses and high-volume rinses is recommended to mitigate the risk of infection [6, 11]. Despite implementing standard safeguards for dentists and staff during dental procedures, exposure to aerosols and particles remains a possibility [12]. Previous studies have suggested a plausible association between aerosol exposure and an elevated risk of respiratory, liver, kidney, and neurological dysfunction [13]. Nevertheless, there exists a dearth of information regarding the relationship between dentist health and aerosol exposure during dental treatments.

revious research has demonstrated a noteworthy increase in the production of particulate matter (PM) by dentists during surgical procedures [14]. Similarly, clinical studies conducted in China have also observed significantly elevated PM levels in the dental working environment, in comparison to non-working states [15]. In order to mitigate aerosol contamination within the dental setting, several clinical measures can be implemented. Firstly, regular disinfection of the dental office environment and equipment is crucial [16]. Secondly, the installation of valves and filters on treatment chairs serves to prevent the negative impact of liquids, aerosols, and PM when utilizing suction devices [17]. However, it is important to note that the efficacy of these measures may be limited, potentially leaving dental professionals and patients vulnerable to the inadvertent transmission of various factors that lack evident symptoms [18]. Furthermore, due to the inability of patients to wear masks during treatment, the development of effective aerosol prevention measures within dental offices presents a significant challenge [19]. Therefore, it is imperative to explore more effective approaches to reduce the presence of bacteria, aerosols, and PM.

Mechanical ventilation and air filtration are regarded as preferable preventive measures for mitigating the risk of airborne diseases within the dental office environment. The Centers for Disease Control and Prevention (CDC) recommends enhancing ventilation systems and incorporating air purifiers in dental offices, thereby minimizing potential risks associated with aerosols [20]. Nonetheless, until recent times, limited attention has been given to the size and concentration of PM produced during dental treatments. Consequently, the objective of this study is to assess the efficacy of two types of air removal devices, specifically forceful suction and air disinfection machines, for the purpose of removing bacteria, PM, and aerosols from the dental office environment.

Materials and methods

Tooth sample collection

The collection of tooth samples for this study was conducted in accordance with the ethical standards set forth by the ethics committee of Hebei Medical University and the institutional review board (IRB: 2,011,322,332). Prior to sample collection, all patients were fully informed of the study’s purpose and procedures and provided written informed consent. The study adhered to the guidelines outlined in the CONSORT statement and was conducted in compliance with the Declaration of Helsinki. Inclusion criteria for the study required poor prognosis and the extraction of teeth under constant pressure. A total of 100 teeth were collected from recruited patients, which were obtained from specimens acquired by professional dentists treating severe periodontal disease or severe caries. All samples had fully or partially filled occlusal surfaces and were rapidly extracted and placed in sterile distilled water.

Experimental model for particle exposure assessment

To simulate a dentist operating in an office setting, we utilized a head shape model based on a previous study [5]. The mobile dental equipment(Zhuhai Duojun Biotechnology Co., Ltd. Zhuhai, China) controlled the drilling and grinding of dental samples, while the high-speed rotating instrument had a speed and efficiency similar to clinical use and produced 50–60 LPM of water to reduce the temperature of the drilling surface [5]. The rotary speed and work efficiency of the high-speed instrument was 350,000–400,000 cycles per minute and 50–60 LPM of water was produced to reduce the temperature of the drilling surface.

The efficiency of forceful suction and air disinfection

The pump delivery system was adjusted to a flow rate of 8.5 L per minute (LPM) to replicate the nasal inhalation of a human wearing a surgical mask. The mask was secured tightly to the face using taped fittings along the edges to ensure a perfect fit. Two air removal systems were employed to evaluate the efficiency of PM removal.

Prior to commencing drilling or grinding, the indoor background particle concentration was measured. The grinding process was performed for a duration of 2 min, and air samples were collected using a central vacuum system. Each of the 100 teeth was drilled using a new grinding drill. The size and concentration of aerosol particles were measured at a distance of 15 cm from the tooth. Of the 100 teeth, 50 were treated with a N95 mask (Winner Cor. Guanzhou, China) only, 25 with a N95 mask (Winner Cor. Guanzhou, China) and forceful suction (Philips Cor. Netherlands), and 25 with a N95 mask (Winner Cor. Guanzhou, China) and air disinfection machines(Philips Cor. Netherlands).

Particle counters (model 1.109, Grimm Labortechnik Ltd., Ainring, Germany) were employed to assess the size of aerosol particles generated during the drilling and grinding of teeth. The detector was capable of detecting aerosol particles ranging from 0.26 to 34 mm and recorded data every 6 s. The concentrations of PM ≥ 0.5 (aerodynamic diameter ≥ 0.5 μm), PM10 (≤ 10 μm), PM2.5 (≤ 2.5 μm), PM2.5–10 (2.5 < da ≤ 10 μm), and PM1 (≤ 1 μm) were recorded. Hand-held nanoanalyzers (NanoFCM Co., Ltd, Lincoln,USA) were used to detect aerosol particles ranging from 10 to 300 nm and to detect changes in nanoparticle concentration.

Room airflow and mechanical ventilation rates

The study measured the volumetric airflow rates of enclosed dental treatment rooms and open bay clinics in cubic feet per minute (CFM or ft3/min), using an air velocity sensor integrated into a flow hood (ADM-850 L Airdata Multimeter with CFM-850 L FlowHood, Shortridge Instruments, Inc., Scottsdale, AZ). For metric unit conversion, 1 CFM equals 0.0283 cubic meters per minute (m3/min). The manufacturer calibrated the flow hood of Airdata multimeter following a program that complies with the ANSI/NCSL Z540-1, ISO 17,025, and MIL-STD 45,662 A standards immediately before the experiments. The volumetric sizes of the dental treatment rooms and open bay clinics were calculated in cubic feet (CF or ft3), based on the length, width, and ceiling height of each space. The mechanical ventilation rates of each space in the number of air changes per hour (ACH) were calculated as in previous studies [21, 22].

Statistical analyses

Statistical analyses were conducted by repeating each experiment five times. SPSS 22.0 software (SPSS, Chicago, Illinois) was used for statistical analyses, with a significance level set at 0.05. GraphPad Prism 6.0 software (GraphPad Software, Inc., USA) was used to graph data. The Wilcoxon signed-rank test was used to determine the removal efficiency of aerosol particles of various sizes (PM ≥ 0.5, PM10, PM2.5, PM1) with or without central vacuum during molar grinding. The Mann-Whitney U test was used to compare differences in the PM levels of drilling teeth and indoor air background, as well as in the filtration efficiency of forceful suction and air disinfection machines.

Results

In this study, we conducted an analysis of the concentration of total particulate matter (PM) before and after teeth grinding. 50 cases in no suction group and 50 cases in suction group. In addition, for suction group, 25 people in the forceful suction group and 25 people in the air disinfection group.

Our findings indicate that the concentration of total PM generated during teeth grinding was significantly higher than the indoor background concentration (p < 0.001, as shown in Fig. 1). When using the drilling and grinding process without vacuum, 0.75% of the aerosol particles in the drilling and grinding process were PM2.5-10, 78.25% of TSP were PM2.5, and 98.68% of TSP were PM1. Furthermore, the nanoanalyzer measurements revealed that the majority of aerosol particles had an aerodynamic diameter below 60 nm, with an average particle diameter of 52.61 nm and an average concentration of 2.6*1011 ultrafine aerosol particles (as shown in Fig. 2).

Fig. 1
figure 1

Particle size distributions of drilling the teeth and non-drilling the teeth in non-suction conditions

Fig. 2
figure 2

Particle size and concentration distributions of ultrafine aerosol particles generated from drilling teeth procedures

We also investigated the impact of air suction devices on airborne aerosol particles. Our results demonstrate that the air changes per hour (ACH) level significantly decreased in the suction devices group (p < 0.001, as shown in Table 1). Additionally, the number concentration of PM ≥ 0.5 and PM2.5 was significantly lower in the suction group compared to the group without suction. Mass concentrations of PM such as PM ≥ 0.5, PM10, PM1.0, and PM2.5 were also significantly lower in the suction group compared to the group without suction (as shown in Table 1).

Table 1 Comparison of particle concentrations under drilling the teeth without/with suction conditions

Furthermore, we analyzed the effects of different air exchange machine, such as forceful suction and air disinfection machines, during teeth grinding. Our analysis revealed that the ACH was significantly lower in the air disinfection machines group compared to the forceful suction group. Additionally, the number and mass concentration of aerosol particles were significantly lower in the air disinfection machines group compared to the forceful suction group in terms of PM2.5 levels. However, the forceful suction group also reduced the mass concentration in PM10 level compared to the air disinfection machines group (p = 0.011, supplement Table 1).

Figure 3 demonstrates that air disinfection machines (DM) were more effective in removing aerosol particles than forceful suction (FS) at 4, 8, 12, 16, 26, and 32 min. Our analysis of different air removal devices revealed that ventilation systems were significantly associated with air disinfection machines, and we correlated the time taken for air changes per hour (ACH) with 95% and 100% aerosol particle removal. The correlation coefficients were high between ACHdm and the time needed to reach 95% (r=-0.92, p < 0.01) and 100% (r=-0.85, p < 0.01) aerosol particle removals with air disinfection machines. Similarly, the correlation was clear between ACH forceful suction and the time needed to reach 95% (r=-0.89, p < 0.05) and 100% (r=-0.75, p < 0.05) aerosol removals with the forceful suction group, as illustrated in Fig. 4.

Fig. 3
figure 3

Removal efficiency for the 0.3 μm aerosol aerosol particles with forceful suction and air disinfection machines at 4, 8, 12, 16, 26 and 32 min after aerosol generations in dental treatment rooms with various mechanical ventilation rates measured by air change per hour (ACH).

Fig. 4
figure 4

Correlations between air disinfection machines (ACHdm) or forceful suction (ACHfs) and times needed to 95% or 100% removals of the 0.3 μm aerosol aerosol particles. Higher ACHdm(A, B) and ACHfs(C,D) are highly correlated with shorter times needed to reach 95% or 100% aerosol removal

Discussion

In the discussion, it is noted that the drilling and grinding of teeth can increase the concentration of suspended aerosol particles in the dental environment. However, other sources can also contribute to higher particle concentrations in dental offices [23]. A recent study in the United States found that indoor aerosol particles in dental offices were primarily composed of aerosol particles 6500 nm in size, with ultrafine aerosol particles (< 100 nm in size) accounting for 67% of all aerosol particles [24]. Our study also found a significant increase in aerosol particles after the start of drilling, and these fine aerosol particles may penetrate deep into the lungs through the alveoli of the respiratory system [25].

Dental droplets or aerosols, generated by the aerodynamic forces resulting from high-speed rotations or ultrasonic vibrations of dental instruments, possess the potential to project into the air within dental treatment rooms, leading to contamination of the indoor environment [11, 26]. The differentiation between droplets and aerosols has traditionally been established based on their physical size, with particles larger than 5 μm in diameter being classified as droplets that exhibit rapid descent to the ground and limited travel within a 2-meter radius. In contrast, particles smaller than 5 μm are referred to as droplet nuclei or aerosols that can remain suspended in the air over an extended duration, traveling distances greater than 2 m [27]. Dental droplets or aerosols, propelled by the aerodynamic forces of high-speed rotation or ultrasonic vibration from dental instruments, can disseminate into the air space of the dental treatment room, leading to contamination of the indoor environment [11, 26]. The distinction between droplets and aerosols is primarily attributed to variations in particle size, where droplets larger than 5 μm in diameter rapidly descend to the ground, covering distances no greater than 2 m. In contrast, droplet nuclei and aerosols smaller than 5 μm can persist in the air for prolonged periods [28]. The smaller the aerosol particles, the longer their airborne suspension, consequently increasing the risk of inhalation by individuals in the vicinity. Although low- and high-volume vacuum evacuation serve as effective measures for controlling droplets and aerosols, their implementation can reduce microbial and particle contamination by more than 90%. However, it is important to acknowledge that the health implications associated with these minute aerosol particles within the dental environment still require further investigation. Nevertheless, the Centers for Disease Control and Prevention (CDC) advises dental professionals to restrict aerosol-generating procedures and implement measures to enhance ventilation and air filtration measures during the ongoing COVID-19 pandemic [29]. S Given the prolonged suspension of aerosol particles within this size range in the air, improving ventilation and utilizing air filtration systems emerge as the most feasible and efficacious approaches to eliminate them from the indoor environment [30, 31]. Aerosol refers to a dispersion system comprising solid or liquid aerosol particles suspended in a gaseous medium, which can be generated by people, animals, instruments, or machines. In December 2019, an outbreak of pneumonia caused by a novel coronavirus infection emerged in China, with respiratory droplet transmission and close contact transmission being the primary modes of transmission [32,33,34]. In relatively closed environments with prolonged exposure to high concentrations of aerosols, the possibility of aerosol transmission cannot be ruled out [35,36,37]. Given that dental treatments generate a large number of droplets and aerosols carrying pathogenic microorganisms, there is a high risk of in-hospital spread and hospital infection [38,39,40].

Furthermore, comprehending the transmission of respiratory pathogens holds paramount importance for implementing effective public health measures aimed at mitigating their spread [41, 42]. The generation and size of particles play pivotal roles in the carriage, atomization, and dissemination of pathogens. The generation of infectious respiratory particles hinges upon the type and frequency of respiratory activities, the nature and site of infection, and the pathogen load [43]. Additionally, relative humidity, particle aggregation, and mucus properties exert influence over the size of expelled particles and their subsequent dispersion [44].

The dental profession possesses unique characteristics, whereby the examination, diagnosis, and treatment procedures are conducted within the confines of the dental office. Within this integrated treatment space, various factors such as doctor-patient communication and patient behavior (e.g., coughing, sneezing, gargling) can generate droplets. Additionally, the use of high-speed ultrasonic equipment during treatment can produce a substantial amount of aerosols and droplets [45,46,47,48]. Droplets that are less than 5 μm in diameter remain suspended in the air, and the nucleus of droplets formed after dehydration can move with the air, thus leading to environmental pollution in the treatment area [44, 49]. Larger diameter droplets tend to settle quickly in the vicinity of the droplet source. If not processed in a timely manner, the settled droplets can form droplet nuclei after drying, which can be mechanically suspended into the air again, resulting in secondary pollution [50, 51]. Therefore, it can be inferred that the primary components of the aerosol in the dental treatment area are the aerosol generated by the treatment process, the droplet nucleus formed after drying the droplet sprayed into the air, and the droplet nucleus raised again after drying and dehydration of the larger diameter droplet settlement.

Oral treatment is a continuous and dynamic process, where the number of chairs in operation during a given period of time directly affects the bacterial content in the air and the settling time required after the treatment. A study revealed that two hours after the ultrasonic scaling operation, the bacterial content in the air increased by 4.8 times in a five-chair office and 4.3 times in a single-chair office. Furthermore, one hour after the treatment, the bacterial colonies in the air decreased by 45% in a five-chair office, which was 2.3 times more than before the opening, and 39% in a single-chair office, which was 1.9 times more than before the opening [52, 53]. The range of bioaerosol contamination varies depending on the operation, with aerosols from ultrasonic scaling spreading up to 1 m horizontally and 0.5 m vertically, and spattering distances up to 1.6 m horizontally and 1.8 m vertically for dental preparation. In a closed room, aerosol contamination is more extensive and can reach the entire room [54,55,56].

There is a clear association between fixed particles and pulmonary complications. First of all, there is a clear correlation between the time solid particles stay in the lungs, and the larger the particles, the longer the lung remains [57]. Second, there is also a clear association between solid particles and the incidence of lung disease [58]. One of the most notable instances of an air pollution catastrophe occurred in London in December 1952, resulting in an estimated count of over 4,000 deaths [59]. Air pollution comprises a intricate amalgamation of gases and particles. Gaseous pollutants infiltrate deeply into the alveoli, facilitating their diffusion across the blood-air barrier to affect numerous organs [60]. On the other hand, particulate matter (PM) encompasses a concoction of solid or liquid particles suspended in the atmosphere. Depending on their size, coarse particles (PM10) settle in the upper airways, whereas fine particles (PM2.5) can accumulate within the lung parenchyma, instigating various respiratory ailments. Besides size, the composition of PM has been linked to diverse toxicological consequences based on clinical, epidemiological, in vivo, and in vitro animal and human studies. PM can consist of organic, inorganic, and biological compounds, all of which possess the capability to alter several biological activities, including cytokine production, coagulation factor equilibrium, pulmonary function, respiratory symptoms, and cardiac function [61, 62]. As it traverses the airways, the exposure to air pollution can engender various alterations, including the recruitment of inflammatory cells and the subsequent release of cytokines and reactive oxygen species (ROS). These inflammatory agents have the capability to activate distinct signaling pathways, such as MAP kinases, NF-κB, and Stat-1, as well as induce DNA adducts [63]. Collectively, these modifications can contribute to the development of obstructive or restrictive pulmonary diseases, encompassing conditions such as asthma, chronic obstructive pulmonary disease (COPD), pulmonary fibrosis, and even cancer [64]. In 2013, based on a comprehensive analysis of research studies pertaining to the effects of air pollution, the International Agency for Research on Cancer (IARC) classified outdoor air pollution as Group 1 [65]. Harrel SK et al. found that aerosol concentrations were high within a 2-foot (0.609 m) range centered on the patient [11, 66]. Zhang Yuqin et al. demonstrated that the degree of contamination in the dental office was inversely proportional to the distance from the treatment site, decreasing with increasing distance [67]. Our study employed two air exchange systems that significantly improved airborne aerosol particles, not only in terms of concentration but also in terms of particle size. This finding is consistent with previous studies that suggest suction devices may reduce aerosol concentrations during dental procedures [68]. Additionally, our results indicated that the air disinfection machines group significantly reduced airborne aerosol particles compared to the forceful suction group.

Limitations of this study include the challenge of accurately quantifying the precise exposure of dental staff. Although we utilized a simulation system to analyze the removal of airborne aerosol particles, it may not fully reflect actual working conditions. Furthermore, other factors such as sound, humidity, and wind speed are crucial in determining particle distribution and require further investigation in future studies.

Conclusion

In conclusion, the air exchange system can effectively decrease the number of aerosol particles generated during dental procedures involving drilling and grinding. Comparing the two air exchange systems, we found that the group utilizing air disinfection machines demonstrated a more significant reduction in suspended aerosol particles compared to the forceful suction group.

Data Availability

All data generated or analysed during this study are included in this published article.

Change history

Abbreviations

PM:

Suspended particles

FS:

Forceful suction

DM:

Air disinfection machines

TSP:

Total suspended particles

CDC:

Centers for Disease Control

LPM:

Liter per minute

ACH:

Air change per hour

ACHdm:

Air disinfection machines

ACHfs:

Forceful suction

References

  1. Selwitz RH, Ismail AI, Pitts NB. Dent Caries Lancet. 2007;369(9555):51–9.

    Article  Google Scholar 

  2. Bostanci N, et al. Periodontal disease: from the lenses of light microscopy to the specs of proteomics and next-generation sequencing. Adv Clin Chem. 2019;93:263–90.

    Article  PubMed  Google Scholar 

  3. Baakdah RA, et al. Pediatric dental treatments with pharmacological and non-pharmacological interventions: a cross-sectional study. BMC Oral Health. 2021;21(1):186.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Nordendahl E, et al. Invasive Dental treatment and risk for a first myocardial infarction. J Dent Res. 2018;97(10):1100–5.

    Article  PubMed  Google Scholar 

  5. Liu M-H et al. Removal efficiency of central vacuum system and protective masks to su spended particles from dental treatment. PLoS ONE 14(11): p. e0225644.

  6. Bentley CD, Burkhart NW, Crawford JJ. Evaluating spatter and aerosol contamination during dental procedures. J Am Dent Assoc. 1994;125(5):579–84.

    Article  PubMed  Google Scholar 

  7. Eisen S, Eisen D, Farmelant J. Evaluation of spatter generation and contamination during instrument cleaning prior to sterilization. J Mass Dent Soc. 2006;55(2):26–9.

    PubMed  Google Scholar 

  8. Grenier D. Quantitative analysis of bacterial aerosols in two different dental clinic environments. Appl Environ Microbiol. 1995;61(8):3165–8.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Grenier D. Quantitative analysis of bacterial aerosols in two different dental cl inic environments. Appl Environ Microbiol. 61(8): p. 3165–8.

  10. Hallier C et al. A pilot study of bioaerosol reduction using an air cleaning system dur ing dental procedures. Br Dent J 209(8): p. E14.

  11. Harrel SK, Molinari J. Aerosols and splatter in dentistry: a brief review of the literature and infection control implications. J Am Dent Assoc. 2004;135(4):429–37.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Bennett AM, et al. Microbial aerosols in general dental practice. Br Dent J. 2000;189(12):664–7.

    Article  PubMed  Google Scholar 

  13. Hong YJ, et al. Assessment of volatile organic compounds and particulate matter in a dental clinic and health risks to clinic personnel. J Environ Sci Health A Tox Hazard Subst Environ Eng. 2015;50(12):1205–14.

    Article  PubMed  Google Scholar 

  14. Sotiriou M, et al. Measurement of particle concentrations in a dental office. Environ Monit Assess. 2008;137(1–3):351–61.

    Article  PubMed  Google Scholar 

  15. Lang A et al. Nanoparticle concentrations and composition in a dental office and den tal laboratory: a pilot study on the influence of working procedures. J Occup Environ Hyg. 15(5): p. 441–7.

  16. Szymańska J. Dental bioaerosol as an occupational hazard in a dentist’s workplace. Ann Agric Environ Med. 2007;14(2):203–7.

    PubMed  Google Scholar 

  17. Polednik B. Aerosol and bioaerosol particles in a dental office. Environ Res. 2014;134:405–9.

    Article  PubMed  Google Scholar 

  18. Cao Y, et al. Enterotoxigenic Bacteroidesfragilis promotes intestinal inflammation and malignancy by inhibiting exosome-packaged miR-149-3. Gastroenterology. 2021;161(5):1552–1566e12.

    Article  PubMed  Google Scholar 

  19. Hu Z et al. Clinical characteristics of 24 asymptomatic infections with COVID-19 s creened among close contacts in Nanjing, China. Sci China Life Sci. 63(5): p. 706–11.

  20. Aljarbou FA, et al. Clinical dental students’ knowledge regarding proper dental settings for treating patient during COVID-19: a cross-sectional study. Pak J Med Sci. 2021;37(2):503–9.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Conducting sputum induction safely %J Examination 1999.

  22. Ren Y-F et al. Effects of mechanical ventilation and portable air cleaner on aerosol removal from dental treatment rooms. J Dent. 105: p. 103576.

  23. Lang A, et al. Nanoparticle concentrations and composition in a dental office and dental laboratory: a pilot study on the influence of working procedures. J Occup Environ Hyg. 2018;15(5):441–7.

    Article  PubMed  Google Scholar 

  24. Wu XM, Apte MG, Bennett DH. Indoor particle levels in small- and medium-sized commercial buildings in California. Environ Sci Technol. 46(22): p. 12355–63.

  25. Darquenne C. Aerosol deposition in health and disease. J Aerosol Med Pulmonary drug Delivery. 25(3): p. 140–7.

  26. Zemouri C, et al. A scoping review on bio-aerosols in healthcare and the dental environment. PLoS ONE. 2017;12(5):e0178007.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Qian H, et al. Particle removal efficiency of the portable HEPA air cleaner in a simulated hospital ward. Build Simul. 2010;3(3):215–24.

    Article  Google Scholar 

  28. Gralton J, et al. The role of particle size in aerosolised pathogen transmission: a review. J Infect. 2011;62(1):1–13.

    Article  PubMed  Google Scholar 

  29. Jayaweera M, et al. Transmission of COVID-19 virus by droplets and aerosols: a critical review on the unresolved dichotomy. Environ Res. 2020;188:109819.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Siegel JD et al. 2007 Guideline for Isolation Precautions: Preventing Transmission of Infectious Agents in Health Care Settings Am J Infect Control, 2007. 35(10 Suppl 2): p. S65-164.

  31. Fennelly KP. Particle sizes of infectious aerosols: implications for infection control. Lancet Respir Med. 2020;8(9):914–24.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Chou R, et al. Epidemiology of and risk factors for coronavirus infection in Health Care Workers: a living Rapid Review. Ann Intern Med. 2020;173(2):120–36.

    Article  PubMed  Google Scholar 

  33. Torales J, et al. The outbreak of COVID-19 coronavirus and its impact on global mental health. Int J Soc Psychiatry. 2020;66(4):317–20.

    Article  PubMed  Google Scholar 

  34. Wen Z, et al. Cytoskeleton-a crucial key in host cell for coronavirus infection. J Mol Cell Biol. 2020;12(12):968–79.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Bao F, et al. Estimating the columnar concentrations of Black Carbon Aerosols in China using MODIS Products. Environ Sci Technol. 2020;54(18):11025–36.

    Article  PubMed  Google Scholar 

  36. Garzona-Navas A, et al. Mitigation of Aerosols Generated during Exercise Testing with a Portable High-Efficiency Particulate Air Filter with Fume Hood. Chest. 2021;160(4):1388–96.

    Article  PubMed  Google Scholar 

  37. Wang Y, et al. Toxicity of Ortho-phthalaldehyde Aerosols in a human in Vitro Airway tissue model. Chem Res Toxicol. 2021;34(3):754–66.

    Article  PubMed  Google Scholar 

  38. Drohan SE, et al. Incentivizing hospital infection control. Proc Natl Acad Sci U S A. 2019;116(13):6221–5.

    Article  PubMed  PubMed Central  Google Scholar 

  39. Emmerson AM. The impact of surveys on hospital infection. J Hosp Infect. 1995;30(Suppl):421–40.

    Article  PubMed  Google Scholar 

  40. Gray J, Atherton N. The Journal of Hospital infection: moving with the times. J Hosp Infect. 2016;92(2):113–4.

    Article  PubMed  Google Scholar 

  41. Johnson GR, Morawska L. The mechanism of breath aerosol formation. J Aerosol Med Pulm Drug Deliv. 2009;22(3):229–37.

    Article  PubMed  Google Scholar 

  42. Garner JS, Simmons BP. Guideline for isolation precautions in hospitals. Infect Control. 1983;4(4 Suppl):245–325.

    PubMed  Google Scholar 

  43. Booth TF, et al. Detection of airborne severe acute respiratory syndrome (SARS) coronavirus and environmental contamination in SARS outbreak units. J Infect Dis. 2005;191(9):1472–7.

    Article  PubMed  Google Scholar 

  44. Xie X, et al. How far droplets can move in indoor environments–revisiting the Wells evaporation-falling curve. Indoor Air. 2007;17(3):211–25.

    Article  PubMed  Google Scholar 

  45. Apelian N, Vergnes JN, Bedos C. Is the dental profession ready for person-centred care? Br Dent J. 2020;229(2):133–7.

    Article  PubMed  Google Scholar 

  46. Luo YL, et al. A survey of dental therapists’ practice patterns and training in Minnesota. J Am Dent Assoc. 2021;152(10):813–21.

    Article  PubMed  Google Scholar 

  47. MacEntee MI. Does the dental profession care for disabled elders? Some practical questions. J Can Dent Assoc. 1990;56(3):215–7.

    PubMed  Google Scholar 

  48. Selwitz RH. The dental profession’s role in programs for detection of high blood pressure. J Public Health Dent. 1977;37(4):253–65.

    Article  PubMed  Google Scholar 

  49. Marze S, Nguyen HT, Marquis M. Manipulating and studying triglyceride droplets in microfluidic devices. Biochimie. 2020;169:88–94.

    Article  PubMed  Google Scholar 

  50. Air cleaning technologies: an evidence-based analysis. Ont Health Technol Assess Ser, 2005. 5(17): p. 1–52.

  51. Khater A, et al. Dynamics of temperature-actuated droplets within microfluidics. Sci Rep. 2019;9(1):3832.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Krishna R, De Stefano JA. Ultrasonic vs. hand instrumentation in periodontal therapy: clinical outcomes. Periodontol 2000. 2016;71(1):113–27.

    Article  PubMed  Google Scholar 

  53. Mannakandath ML, et al. Effect of ultrasonic scaling with adjunctive photodynamic therapy on the treatment of gingival inflammation among diabetic patients undergoing fixed orthodontic treatment. Photodiagnosis Photodyn Ther. 2021;35:102360.

    Article  PubMed  Google Scholar 

  54. Fink JB, et al. Reducing aerosol-related risk of transmission in the era of COVID-19: an interim Guidance endorsed by the International Society of Aerosols in Medicine. J Aerosol Med Pulm Drug Deliv. 2020;33(6):300–4.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Li J, Fink JB, Ehrmann S. High-flow nasal cannula for COVID-19 patients: low risk of bio-aerosol dispersion. Eur Respir J, 2020. 55(5).

  56. Mirskaya E, Agranovski IE. Sources and mechanisms of bioaerosol generation in occupational environments. Crit Rev Microbiol. 2018;44(6):739–58.

    Article  PubMed  Google Scholar 

  57. Huang Z, et al. Relationship between particle size and lung retention time of intact solid lipid nanoparticle suspensions after pulmonary delivery. J Control Release. 2020;325:206–22.

    Article  PubMed  Google Scholar 

  58. Jie Y, et al. Relationship between pulmonary function and indoor air pollution from coal combustion among adult residents in an inner-city area of southwest China. Braz J Med Biol Res. 2014;47(11):982–9.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Osornio-Vargas AR, et al. In vitro biological effects of airborne PM2.5 and PM10 from a semi-desert city on the Mexico-US border. Chemosphere. 2011;83(4):618–26.

    Article  PubMed  Google Scholar 

  60. Verma V, et al. Contribution of water-soluble and insoluble components and their hydrophobic/hydrophilic subfractions to the reactive oxygen species-generating potential of fine ambient aerosols. Environ Sci Technol. 2012;46(20):11384–92.

    Article  PubMed  Google Scholar 

  61. Betha R, Balasubramanian R. Emissions of particulate-bound elements from biodiesel and ultra low sulfur diesel: size distribution and risk assessment. Chemosphere. 2013;90(3):1005–15.

    Article  PubMed  Google Scholar 

  62. Khalek IA, et al. Regulated and unregulated emissions from modern 2010 emissions-compliant heavy-duty on-highway diesel engines. J Air Waste Manag Assoc. 2015;65(8):987–1001.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Wang D, et al. Macrophage reactive oxygen species activity of water-soluble and water-insoluble fractions of ambient coarse, PM2.5 and ultrafine particulate matter (PM) in Los Angeles. Atmos Environ. 2013;77:301–10.

    Article  Google Scholar 

  64. Ghio AJ, et al. Metals associated with both the water-soluble and insoluble fractions of an ambient air pollution particle catalyze an oxidative stress. Inhal Toxicol. 1999;11(1):37–49.

    Article  PubMed  Google Scholar 

  65. Falcon-Rodriguez CI, et al. Aeroparticles, composition, and Lung Diseases. Front Immunol. 2016;7:3.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Harrel SK, Barnes JB, Rivera-Hidalgo F. Aerosol and splatter contamination from the operative site during ultrasonic scaling. J Am Dent Assoc. 1998;129(9):1241–9.

    Article  PubMed  Google Scholar 

  67. Xiao X et al. Characterization of Odontogenic Differentiation from Human Dental Pulp Stem Cells Using TMT-Based Proteomic Analysis Biomed Res Int, 2020. 2020: p. 3871496.

  68. Rupf S et al. Exposure of patient and dental staff to fine and ultrafine particles from scanning spray. Clin Oral Invest. 19(4): p. 823–30.

Download references

Acknowledgements

We would like to thank all participants and our hospital.

Funding

This research was partile funded by Medical Science Research Project of Hebei Province (No.20210048) and Hebei Provincial Government for the Training of Outstanding Clinical Medical Talents and Basic Research Project (No. 361029 (MXZB00271)).

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: Yaru Du, Fei Zhao, Data curation: Ran Tao, Bing Liu, Formal analysis: Ran Tao, Bing Liu, Funding acquisition: Fei Zhao, Ran Tao, Bing Liu, Investigation: Fei Zhao, Ran Tao, Bing Liu, Methodology: Fei Zhao, Ran Tao, Bing Liu, Project administration: Fei Zhao, Ran Tao, Bing Liu, Resources: Fei Zhao, Ran Tao, Bing Liu, Software: Fei Zhao, Supervision: Fei Zhao, Validation: Fei Zhao, Visualization: Fei Zhao, Writing – original draft: Fei Zhao, Ran Tao, Bing Liu, Writing – review & editing: Yaru Du, Fei Zhao, Ran Tao, Bing Liu.

Corresponding author

Correspondence to Bing Liu.

Ethics declarations

Ethics approval and consent to participate

This study was approved and approved by the ethics committee of Hebei Medical University and the institutional review board (IRB: 2011322332). All patients were informed and signed an informed consent form. According to the CONSORT guidelines, these studies also comply with the Declaration of Helsinki.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Du, Y., Zhao, F., Tao, R. et al. Effect of forceful suction and air disinfection machines on aerosol removal. BMC Oral Health 23, 652 (2023). https://doi.org/10.1186/s12903-023-03369-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12903-023-03369-1

Keywords