Fact-checked by Grok 2 weeks ago

Ultraviolet germicidal irradiation

Ultraviolet germicidal irradiation (UVGI) is a disinfection technology that utilizes ultraviolet-C (UV-C) light, typically at a wavelength of 254 nm emitted by low-pressure mercury lamps, to inactivate microorganisms such as bacteria, viruses, and fungi by damaging their genetic material through the formation of pyrimidine dimers in nucleic acids (DNA and RNA), thereby preventing replication and rendering them harmless. This method has been employed for over a century to purify air, water, and surfaces in controlled environments, offering a chemical-free alternative to traditional sanitization techniques. The foundational principles of UVGI trace back to late 19th-century observations of sunlight's effects, with pioneering work in by William F. Wells demonstrating its potential for controlling airborne infections like through upper-room irradiation systems. Interest in UVGI declined mid-century with the rise of antibiotics but revived in the 1980s amid resurgent outbreaks and later expanded during the for its efficacy against and other respiratory pathogens. UVGI systems are categorized into upper-room configurations, where fixtures mounted high in occupied spaces irradiate air above head level with mechanical mixing to circulate pathogens into the treatment zone; in-duct installations within HVAC systems for whole-building air treatment; and portable or full-room units for unoccupied areas. These applications are prevalent in healthcare facilities, schools, offices, and shelters to supplement and , potentially achieving equivalent air cleaning rates to multiple outdoor air changes while conserving energy. Effectiveness of UVGI depends on factors such as dose, relative humidity (optimal below 60–70%), and air circulation, with historical studies showing reductions in transmission from 34.9% to 9.5% in treated wards and incidence dropping from 53.6% to 13.3% in irradiated schools. Modern research confirms its potency against a broad spectrum of pathogens, including and coronaviruses, often outperforming increased alone in reducing airborne risk. Safety protocols are essential, as direct UV-C exposure can cause skin and ; thus, systems require shielding, annual lamp replacements, and trained maintenance to protect occupants.

Principles of Operation

Mechanism of UV Inactivation

Ultraviolet germicidal irradiation (UVGI), particularly in the UV-C range of 200–280 nm, inactivates microorganisms primarily through direct absorption of photons by their nucleic acids, with peak absorption occurring at approximately 260 nm due to the π-electron systems in pyrimidine and purine bases. This absorption excites the bases, leading to photochemical reactions that distort the DNA or RNA structure without requiring intermediary molecules. In contrast to longer wavelengths like UV-A or UV-B, where indirect damage via reactive oxygen species (ROS) plays a significant role, UV-C inactivation is predominantly direct, as the high-energy photons (shorter wavelengths) are efficiently absorbed by genetic material, minimizing ROS involvement. The primary form of damage induced by UV-C is the creation of photoproducts in DNA and RNA, most notably cyclobutane pyrimidine dimers (CPDs) and pyrimidine (6-4) pyrimidone photoproducts (6-4 PPs). CPDs, which account for about 75% of lesions, form a covalent four-membered ring between the 5,6-double bonds of adjacent pyrimidine bases (such as thymine-thymine or cytosine-cytosine), while 6-4 PPs involve a bond between the 6-position of the 5'-pyrimidine and the 4-position of the 3'-pyrimidine. These dimers block DNA replication and transcription by introducing helical distortions and preventing the action of DNA and RNA polymerases, ultimately leading to cell death or inability to reproduce. Less commonly, UV-C can cause single-strand breaks or base modifications, but dimer formation remains the dominant mechanism for microbial inactivation. Microorganisms possess repair mechanisms to counteract UV-induced damage, including photoreactivation and (NER), which can restore viability if exposure is insufficient. Photoreactivation involves photolyase enzymes that use visible light (near-UV or ) to directly reverse CPDs by splitting the dimer bonds through , effectively monomerizing the bases without altering the genetic sequence. NER, on the other hand, recognizes distorted sites, excises a short segment containing the (typically 12–13 in ), and resynthesizes the gap using the intact strand as a template. High UV doses overcome these repairs by overwhelming the cellular capacity: excessive dimers saturate repair enzymes, leading to unrepaired lesions that accumulate and cause lethal mutations or replication failure during subsequent . The rate of microbial inactivation by UVGI follows the Chick-Watson model, which assumes first-order kinetics where the logarithm of the surviving population decreases linearly with applied UV dose. To derive the inactivation rate constant k, start from Chick's law for disinfection: the differential equation \frac{dN}{dt} = -k' N (where N is the number of viable microbes and k' is the time-based rate constant) integrates to \ln(N/N_0) = -k' t, or equivalently \log_{10}(N_0/N) = \frac{k' t}{\ln(10)}. For UV, the dose D (in mJ/cm²) replaces time via D = I \cdot t ( I times exposure time t), and since k' is proportional to I, the model simplifies to \log_{10}(N_0/N) = k D, where k (in cm²/mJ) is the dose-based inactivation rate constant. Solving for k yields k = \frac{\log_{10}(N_0/N)}{D}, which quantifies susceptibility: higher k indicates faster inactivation per unit dose. This linear relationship holds for doses achieving multi-log reductions before repair or tailing effects dominate.

Germicidal UV Wavelengths

Ultraviolet radiation is classified into bands based on wavelength: UV-A (315–400 nm), which has minimal germicidal activity due to low absorption by microbial components; UV-B (280–315 nm), which exhibits moderate germicidal effects through partial absorption by nucleic acids and proteins; and UV-C (100–280 nm), the primary band for germicidal applications owing to its strong absorption by DNA and RNA in microorganisms. The peak germicidal action occurs around 260–265 nm, aligning with the maximum absorption peak of DNA, where photons are efficiently absorbed to induce photochemical damage such as thymine dimers. Low-pressure mercury lamps, commonly emitting at 254 nm, provide near-optimal efficacy, achieving approximately 85% of the peak DNA absorption efficiency. Within the UV-C band, far-UVC wavelengths (200–235 nm) offer effective inactivation while posing lower risks to human tissues, as they are predominantly absorbed in the outer dead layers of skin and the tear film of eyes without penetrating to living cells. Atmospheric absorption imposes practical limits on UV-C utilization: wavelengths below approximately 200 nm are strongly absorbed by molecular oxygen (O₂), rendering them ineffective for transmission through air in open environments, while the stratospheric absorbs much of the spectrum from 200–300 nm, preventing significant natural solar exposure but enabling controlled artificial sources for disinfection. As an alternative to the 254 nm emission from mercury lamps, krypton-chloride (KrCl) lamps produce narrow-band output at 222 nm, delivering comparable germicidal performance through enhanced protein damage alongside effects. At 254 nm, DNA exhibits a higher molar absorption coefficient (approximately 10,000–12,000 M⁻¹ cm⁻¹ for ) compared to proteins (around 1,000–5,000 M⁻¹ cm⁻¹, depending on composition), prioritizing targeting over protein denaturation and thus optimizing microbial inactivation with minimal off-target absorption.

UV Dose Calculation

UV dose, also known as fluence, quantifies the cumulative delivered to a surface or medium for germicidal purposes, expressed in units of millijoules per square centimeter (mJ/cm²) or microjoules per square centimeter (μJ/cm²) as standardized by the Ultraviolet Association (IUVA). It is fundamentally calculated as the product of (I, typically in mW/cm²) and exposure time (t, in seconds), yielding the dose D according to : D = I \times t This relationship assumes uniform exposure and represents the ideal energy transfer, with IUVA guidelines specifying mJ/cm² as the primary unit for disinfection applications, such as a target of 40 mJ/cm² to achieve 4-log inactivation of pathogenic viruses in water treatment systems. Measurement of UV dose relies on physical and biological techniques to ensure accuracy in both laboratory and operational settings. Radiometers, calibrated to NIST-traceable standards with uncertainties below 10%, directly measure irradiance at germicidal wavelengths (e.g., 254 nm) and integrate over time to compute dose, providing real-time data for system monitoring. Biodosimetry complements this by employing challenge organisms like MS2 coliphage, a UV-resistant bacteriophage, to validate effective dose delivery through observed log inactivation compared to collimated beam benchmarks; for instance, MS2 dose-response curves enable calculation of reduction equivalent dose (RED) via inactivation kinetics. In practical delivery, UV dose is influenced by system geometry and operational parameters, particularly in dynamic environments. setups approximate parallel rays for uniform exposure in bench-scale tests, minimizing errors to achieve path lengths with less than 10% variation, whereas non-collimated configurations in reactors introduce geometric inefficiencies due to lamp spacing and losses. rates in continuous systems inversely affect , reducing effective exposure as throughput increases (e.g., from 0.94 to 20 million gallons per day), necessitating design adjustments to maintain dose uniformity. For continuous flow reactors, dose calculation incorporates efficiency factors to account for real-world deviations from ideal conditions, expressed as D = I × t × EF, where EF represents the product of lamp output efficiency (declining 20-40% over lamp life due to aging) and design efficiency (influenced by and UV transmittance). Lamp output is typically validated at 100% initial power with fouling factors of 0.4-0.9 applied for sleeve accumulation and aging, while design uses to optimize light distribution and ensure RED delivery exceeds required thresholds under varying flows.

History

Early Discoveries of UV Effects

The foundational observations of (UV) radiation's antimicrobial effects began in the late with experiments on 's inhibitory action on microorganisms. In 1877 and 1878, British researchers Arthur Downes and Thomas P. Blunt demonstrated that exposing test tubes containing solutions, such as Pasteur's , to sunlight prevented , while unexposed controls became turbid with microbial proliferation. Their work showed that the germicidal effect increased with exposure duration and light intensity, and was primarily due to the shorter, violet-blue wavelengths of sunlight, laying the groundwork for understanding UV's selective bactericidal properties. Building on these findings, investigations in the 1880s and 1890s linked UV exposure to therapeutic effects on bacterial skin infections. Danish physician Niels Ryberg Finsen pioneered the use of concentrated UV light from carbon arc lamps to treat , a disfiguring form of cutaneous caused by , achieving notable success by 1895 without causing excessive skin damage. Finsen's systematic studies emphasized UV's ability to penetrate and destroy bacteria in tissues, earning him the 1903 in Physiology or Medicine for introducing phototherapy as a medical tool. Concurrently, in 1890, German bacteriologist reported that cultures of the tubercle bacillus exposed to direct sunlight were rapidly killed, whereas those kept in shade survived, providing direct evidence of sunlight's lethal impact on this and reinforcing UV's role in disinfection. A practical milestone came in 1910 with the installation of the first large-scale UV water disinfection system in , , patented and implemented by researchers including , which treated up to 1 million liters of per day using mercury lamps to generate germicidal UV radiation. This marked an early engineered application of UV principles beyond sunlight. The shift from natural sunlight to artificial sources accelerated around this time, with carbon lamps—initially developed for Finsen's therapies—adapted for lab and disinfection experiments due to their emission of UV-rich spectra, enabling controlled and reliable germicidal exposures independent of weather.

Development in Air Disinfection

In the 1930s, pioneering experiments by William F. Wells at Harvard University demonstrated the potential of ultraviolet germicidal irradiation (UVGI) to inactivate airborne pathogens, particularly focusing on tuberculosis bacteria. Wells and his collaborators showed that UV light at 254 nm wavelength could rapidly kill Mycobacterium tuberculosis suspended in air streams, building on the basic principle of UV-induced DNA damage in microorganisms. These laboratory studies established UVGI as a viable method for air disinfection, shifting attention from surface treatment to controlling airborne transmission via droplet nuclei. From the mid-1930s through the 1940s, UVGI systems were installed in practical settings such as and hospitals to combat respiratory infections, including . Early implementations included wall- or ceiling-mounted UV lamps in pediatric wards and operating rooms, where they reduced airborne bacterial counts by irradiating room air without direct exposure to occupants. In , Wells oversaw the deployment of UVGI in suburban day schools between 1937 and 1941, where upper-room systems—lamps positioned high above occupied spaces to treat circulating air via —prevented the epidemic spread of among children, achieving near-complete control of . The upper-room UVGI concept, first systematically applied by Wells in 1938, became a cornerstone of these developments, allowing safe operation in occupied areas by confining irradiation to the upper portion of rooms while relying on natural air mixing to draw pathogens into the treated zone. Despite these successes, early UVGI systems faced significant challenges, notably dust accumulation on lamp surfaces, which reduced UV output by up to 50% over time and necessitated regular protocols. Researchers developed cleaning schedules and protective shielding to mitigate soiling, ensuring sustained efficacy in dust-prone environments like and hospitals. These practical hurdles informed later designs, emphasizing the need for robust in air disinfection applications.

Development in Water Treatment

The development of ultraviolet germicidal irradiation (UVGI) for water treatment began in the early 20th century, with the first full-scale municipal installation in Marseille, France, in 1910, where UV lamps were used to disinfect drinking water supplied to approximately 300,000 residents. This pioneering plant utilized low-pressure mercury arc lamps submerged in contact chambers, marking the initial practical application of UV for large-scale water purification despite challenges with lamp reliability and water turbidity. Early U.S. trials followed in the 1910s and 1920s, including experimental installations for military camps and small communities, though widespread adoption was limited by the preference for chlorination. Interest in UVGI waned after the 1930s due to the dominance of chemical disinfectants but revived in the and amid growing toward antibiotics and concerns over emerging microbial resistances, prompting renewed into physical disinfection methods. By the , installations expanded in for commercial and residential applications. These mid-century efforts focused on improving reactor designs, transitioning from simple open contact chambers—where water flowed around exposed lamps—to more efficient enclosed systems that minimized recontamination and enhanced UV exposure. From the 1980s onward, the U.S. Environmental Protection Agency (EPA) validated UVGI through rigorous testing protocols, recognizing its efficacy against chlorine-resistant pathogens and publishing guidance on wastewater applications. In the 1990s, standards emerged for inactivating protozoa such as and , with UV doses validated to achieve multi-log reductions in pilot and full-scale studies, driven by outbreaks highlighting chlorination's limitations. A key regulatory milestone came with the 2000 proposal leading to the Long Term 2 Enhanced Treatment Rule (LT2ESWTR), which in 2006 formalized UV as a mandated to chlorination for protozoan inactivation in public water systems, requiring validated reactors to deliver specified UV doses for compliance. Early reactor evolution included straight-tube lamps in parallel arrays within chambers, contrasting with modern coiled or high-output configurations that optimize hydraulic flow and for larger capacities.

Modern Advancements and Revivals

In the 2000s, the emergence of ultraviolet-C (UVC) light-emitting diodes (LEDs) marked a significant advancement in UVGI technology, enabling the development of compact, portable disinfection devices that offered advantages over traditional mercury lamps, such as lower energy consumption and absence of hazardous materials. These LEDs, based on aluminum gallium nitride semiconductors, facilitated on-demand, point-of-care applications in healthcare and consumer settings, with decontamination efficacy comparable to or exceeding conventional systems in controlled tests. By the 2010s, research into far-UVC wavelengths (around 222 nm) gained momentum, demonstrating potential for continuous, safe room irradiation without significant harm to human skin or eyes due to limited tissue penetration. This innovation shifted UVGI toward occupant-present applications, with early studies confirming high pathogen inactivation rates in real-world environments like chambers simulating indoor spaces. The from 2020 to 2023 accelerated UVGI research and deployment, particularly through studies on inactivation. A landmark 2020 trial by researchers showed that far-UVC light at 222 nm achieved 99.9% inactivation of airborne human coronaviruses, including surrogates for , at low doses under realistic indoor conditions. This work built on prior far-UVC evidence, highlighting its efficacy against aerosolized viruses while maintaining safety for continuous exposure in occupied areas. Globally, adoption surged; in 2020, Chinese cities implemented UVC systems in subways and public transit to disinfect surfaces and air, reducing viral contamination in high-traffic environments as part of early pandemic response measures. From 2022 to 2025, UVGI with high-efficiency particulate air () filters in (HVAC) systems became a prominent strategy for enhanced , combining mechanical with germicidal to achieve superior removal in and healthcare settings. The International Ultraviolet Association (IUVA) issued guidelines during this period to support UVGI in pandemic preparedness, emphasizing standardized protocols for efficacy measurement, deployment in HVAC, and safe upper-room applications to mitigate . In 2021, the U.S. granted emergency use authorizations for select UVGI devices in healthcare, facilitating their rapid for surface and air disinfection amid supply shortages. However, far-UVC commercialization faced ongoing challenges by 2025, including low wall-plug in LEDs (typically under 10%) and high costs, limiting widespread adoption despite promising pilot deployments.

Effectiveness

Inactivation Kinetics of Microorganisms

The inactivation of microorganisms by ultraviolet germicidal irradiation (UVGI) typically follows log-linear dose-response curves under ideal conditions, where the logarithm of the surviving population decreases proportionally with the applied UV dose at 254 nm. This relationship is described by first-order kinetics, with the dose required for 90% inactivation (D90) serving as a key metric for susceptibility. D90 values can vary by 2-10x depending on strain, suspension medium (air, water, surface), and detection method (e.g., plaque assay vs. qPCR), with air typically requiring lower doses than turbid water. For example, exhibits a D90 of approximately 3 mJ/cm² in liquid suspensions exposed to low-pressure mercury lamps emitting at 254 nm. In contrast, the bacteriophage MS2, a common viral surrogate, requires a higher D90 of about 2 mJ/cm² in form, reflecting greater resistance due to its single-stranded genome. Resistant organisms, such as protozoan oocysts like Cryptosporidium parvum, often display multi-hit kinetics, where inactivation deviates from strict log-linearity and requires significantly higher doses to achieve substantial reductions. For C. parvum, the D90 is estimated at 1-2 mJ/cm² for monochromatic 254 nm UV in water, necessitating multi-target models to account for the organism's robust oocyst wall and repair mechanisms. UVGI has proven effective against enveloped viruses like SARS-CoV-2, with a D90 of approximately 6.6 J/m² (0.66 mJ/cm²) reported in early 2020 studies using 254 nm irradiation on viral suspensions. However, UVGI is less effective against non-nucleic acid pathogens like prions, which resist inactivation even at high doses due to their proteinaceous nature lacking DNA or RNA targets. Additionally, shadowed or aggregated cells remain protected from direct irradiation, limiting overall efficacy in non-ideal exposures. The Chick-Watson model extends this framework to describe inactivation kinetics, particularly for UVGI, as: \log\left(\frac{N_0}{N}\right) = k \cdot D^n where N_0 and N are initial and final microbial concentrations, D is the UV dose (in mJ/cm²), k is the rate constant (in cm²/mJ), and n is the dilution coefficient (typically ≈1 for most and DNA viruses, indicating behavior). For RNA viruses, n may deviate from 1 (often <1), reflecting sigmoidal curves due to capsid shielding or repair variability. Kinetics vary by organism type: bacteria like E. coli are generally susceptible due to DNA thymine dimer formation, though post-exposure photoreactivation can limit long-term inactivation if repair enzymes are active. Viruses show type-specific resistance, with RNA viruses (e.g., MS2) often more resilient than DNA viruses because RNA bases form fewer lethal photoproducts and lack the same repair constraints. Protozoa and fungi may require higher doses owing to protective structures, while enveloped viruses like SARS-CoV-2 are inactivated efficiently via genome damage but can exhibit tailing at low populations.

Factors Affecting Efficacy

The efficacy of ultraviolet germicidal irradiation (UVGI) in practical applications is influenced by several environmental and operational factors that deviate from ideal laboratory conditions, where baseline inactivation rates are established under controlled exposures. In real-world settings, such as air or water systems, these variables can substantially alter the required UV dose for achieving microbial inactivation, often necessitating system adjustments to maintain performance. Shielding effects from particulate matter, dirt, and biofilms significantly reduce UVGI effectiveness by preventing light penetration to target microorganisms. Suspended particles in water or air can associate with microbes, shielding them and increasing the required UV dose by factors of 2 to 10 times compared to free-floating organisms, as demonstrated in studies on spore-clay aggregates and bacterial flocs. Biofilms, in particular, form protective matrices that scatter UV radiation and harbor aggregated cells, further elevating the dose needed for penetration and inactivation. Similarly, microbial clumping in aerosols or liquids creates internal shadows, reducing overall germicidal impact unless disrupted prior to exposure. Environmental conditions like relative humidity in air and turbidity in water also impair UVGI performance through absorption and scattering mechanisms. In air disinfection systems, relative humidity above 50% can degrade efficacy, with studies showing significant reductions—up to 50% or more—when humidity rises to 75-90%, as water vapor absorbs UV-C wavelengths and forms a protective layer around airborne pathogens. For water treatment, turbidity from suspended solids scatters UV light, diminishing fluence rates and requiring higher doses; even low levels (e.g., 1-5 NTU) can halve the inactivation efficiency of bacteria like E. coli. Organic matter, such as humic acids, exacerbates this by strongly absorbing UV at 254 nm, leading to efficacy drops of approximately 30% in natural waters without pre-treatment, as observed in surface water disinfection experiments. Post-irradiation revival through photoreactivation poses another challenge in light-exposed environments, where UV-damaged DNA in microorganisms like E. coli can be repaired using visible light (300-500 nm), partially reversing inactivation. This process is particularly relevant in water distribution systems or sunlit air pathways, where up to 1-2 log units of bacterial recovery have been reported after initial UV doses, necessitating strategies to minimize subsequent light exposure. To mitigate these factors and enhance UVGI efficacy, pre-filtration is commonly employed to remove particulates, organic matter, and turbidity, improving UV transmittance by 20-50% in pretreated waters. Pulsed UV systems, delivering high-intensity bursts rather than continuous exposure, offer better penetration through aggregates and reduced shielding effects, achieving higher inactivation rates (e.g., 1-2 log additional reduction) against resistant biofilms compared to steady-state irradiation. These enhancements, when integrated, can restore performance closer to laboratory baselines while addressing site-specific limitations.

Safety and Standards

Human Health Risks

Exposure to ultraviolet germicidal irradiation (UVGI), primarily in the UV-C spectrum (100–280 nm), presents acute risks to the eyes, including photokeratitis, commonly referred to as "welder's flash." This condition involves painful inflammation of the cornea and conjunctiva, manifesting as a gritty sensation, excessive tearing, redness, and temporary vision impairment, with symptoms typically onsetting 6–12 hours after exposure and resolving within 24–48 hours. UV-B radiation (280–315 nm), which may be emitted by some UVGI sources, can similarly induce ocular erythema and contribute to photokeratitis through absorption in the corneal layers. Skin exposure to UVGI also carries direct physiological hazards. Acute exposure causes erythema, a reddening and inflammation akin to sunburn, followed by potential blistering or burns if doses are high, due to damage to epidermal cells and vascular dilation. Chronic low-dose exposure heightens the risk of photocarcinogenesis, as UV radiation acts as a complete carcinogen by inducing DNA mutations, suppressing immune responses, and promoting tumor initiation in skin cells. Far-UVC light at 222 nm offers a lower-risk alternative within UVGI, as it is strongly absorbed by proteins in the dead stratum corneum of the skin and the tear film of the eyes, preventing penetration to viable tissues. Clinical studies in 2020 demonstrated its safety for human exposure, with no observed erythema or ocular damage at doses up to 500 mJ/cm² on skin and updated regulatory limits as of 2022 of 161 mJ/cm² for eyes and 479 mJ/cm² for skin permitting continuous exposure over 8 hours without adverse effects. Long-term studies as of 2024, including 36-month evaluations in occupied spaces, have confirmed no acute or chronic ocular or skin damage from continuous exposure within guideline limits. Certain populations face amplified risks from UVGI exposure. Infants possess thinner, more permeable skin with immature protective barriers, rendering them highly susceptible to both acute burns and long-term damage. Fair-skinned individuals, with reduced melanin—a natural UV absorber—are prone to more severe erythema, burns, and elevated carcinogenesis compared to those with darker pigmentation. Historical applications of UV light in 1930s school settings, including therapeutic exposures and early disinfection trials, led to cases of temporary blindness from photokeratitis when protective measures like goggles were absent, underscoring the need for shielding even in controlled environments.

Material Degradation and Environmental Impacts

Ultraviolet germicidal irradiation (UVGI) can induce degradation in various materials, particularly polymers, due to the absorption of UVC photons leading to photochemical reactions such as chain scission and cross-linking. For instance, polyvinyl chloride (PVC) exposed to 254 nm radiation exhibits surface alterations including yellowing and loss of transparency, with noticeable changes occurring after prolonged exposure equivalent to thousands of joules per square centimeter in controlled tests. Similarly, polycarbonate and high-density polyethylene (HDPE) suffer from yellowing, cracking, and reduced tensile strength, with polycarbonate showing up to 47% loss in stress at break after 216 hours of exposure at intensities around 88 µW/cm². These effects are more pronounced in materials with high UV absorption, such as certain plastics and paints, potentially leading to brittleness and accelerated aging in building interiors or equipment components. To mitigate polymer damage, UV-resistant coatings or shielding with reflective materials like aluminum can be employed, while regular inspection and material selection based on reflectivity minimize long-term impacts. In indoor environments, UVGI influences air chemistry by generating reactive species that interact with volatile organic compounds (VOCs). Lamps emitting below 185 nm, such as certain low-pressure mercury variants, produce ozone through oxygen photolysis, which can reach levels exceeding safety thresholds if not filtered. At 254 nm, UVGI promotes the formation of ultrafine particles and oxygenated VOCs via reactions with indoor surfaces and airborne pollutants, with particle concentrations rising from baseline levels to over 10^5 cm⁻³ in experimental chambers. A 2023 study on germicidal UV sources demonstrated that upper-room UVGI elevates hydroxyl radical (OH) concentrations, enhancing oxidation of pollutants like VOCs but also risking the production of irritants such as formaldehyde and secondary aerosols. Ozone mitigation strategies include using doped quartz bulbs to block sub-200 nm emissions or integrating catalytic filters to decompose O₃ post-generation. Environmentally, UVGI systems in large-scale applications contribute to energy consumption, though typically less than alternatives like increased mechanical ventilation. Upper-room systems, for example, add only 0.6% to annual building energy costs in critical spaces, avoiding the 24.5% rise associated with higher outdoor air intake and its attendant greenhouse gas emissions. In wastewater treatment, UVGI produces minimal byproducts compared to chlorination; it avoids significant formation of bromate or trihalomethanes, with nitrite yields limited to about 1% from nitrate under medium-pressure conditions, far below levels from chemical disinfectants. Quartz sleeves in water treatment units protect lamps from fouling while transmitting UVC, reducing maintenance energy and extending system life without introducing additional environmental burdens. Overall, these impacts underscore the need for site-specific assessments to balance disinfection efficacy with material and ecological considerations.

Exposure Limits and Guidelines

The American Conference of Governmental Industrial Hygienists (ACGIH) establishes Threshold Limit Values (TLVs) for occupational exposure to ultraviolet radiation to protect against photochemical effects on eyes and skin. For conventional UV-C at 254 nm, the TLV is 6 mJ/cm² over an 8-hour period for both eyes and skin. In a 2022 update, ACGIH raised the TLV for far-UVC at 222 nm to 161 mJ/cm² for eyes and 479 mJ/cm² for skin over 8 hours, reflecting reduced penetration into sensitive tissues. The International Commission on Non-Ionizing Radiation Protection (ICNIRP) provides guidelines for broad-spectrum UV exposure, recommending that the effective radiant exposure to unprotected skin not exceed 30 J/m² (equivalent to 3 mJ/cm²) over an 8-hour period for wavelengths between 180 nm and 400 nm. These limits are weighted by a spectral effectiveness function to account for varying biological impacts across wavelengths. The U.S. Occupational Safety and Health Administration (OSHA) does not mandate specific permissible exposure limits for UV radiation but aligns its technical guidance with ACGIH TLVs for workplace safety assessments. Far-UVC wavelengths like 222 nm permit 3-10 times higher exposure limits compared to 254 nm due to strong absorption by the corneal protein layer, preventing damage to underlying epithelial cells. Compliance with these limits is validated using biodosimetry, which employs microbial bioindicators to confirm delivered UV doses in systems, alongside direct radiometric measurements. The International Ultraviolet Association (IUVA) supports standards for UV-LED systems, including the 2022 ANSI/IES/IUVA LM-92-22 method for optical and electrical characterization. For high-intensity UVGI applications, guidelines recommend zoning strategies such as restricted access to irradiated areas, lockable controls, and signage to limit unintended human exposure, ensuring operation only by trained personnel in unoccupied or shielded spaces.

Applications

Air Disinfection Systems

Ultraviolet germicidal irradiation (UVGI) is widely implemented in air disinfection systems to mitigate airborne pathogens in indoor environments, particularly through integration with ventilation infrastructure. These systems leverage UV-C light to inactivate microorganisms in the air without direct exposure to occupants, enhancing indoor air quality in settings such as hospitals, schools, and public transit. Common configurations include upper-room, in-duct, and portable units, each designed to achieve effective pathogen reduction while maintaining safety and airflow dynamics. Upper-room UVGI systems position lamps in the upper portion of rooms, shielded by baffles or louvers to direct UV-C light away from occupied spaces and prevent direct human exposure. Air is drawn into this irradiated zone through natural convection or mechanical mixing, where pathogens are inactivated as they pass through the UV field. These systems can achieve disinfection rates equivalent to 10-20 air changes per hour in well-mixed rooms, providing substantial reduction in airborne infectious agents without significantly increasing energy use for ventilation. In-duct UVGI integrates lamps or coils downstream of HVAC filters to treat air as it circulates through ductwork, ensuring broad coverage in centralized systems. Placed post-filters to target captured and airborne microbes, these installations deliver UV doses typically ranging from 10-50 mJ/cm², sufficient for high-log reductions of viruses and bacteria in moving air streams. This approach is particularly effective in large buildings, where it complements filtration by continuously disinfecting recirculated air. The application of UVGI in air disinfection gained renewed emphasis during the COVID-19 pandemic, with the U.S. Centers for Disease Control and Prevention (CDC) issuing guidelines in 2021 recommending upper-room and in-duct systems as supplementary measures to reduce SARS-CoV-2 transmission in indoor settings. These guidelines highlight UVGI's role in providing equivalent clean air changes when combined with ventilation, especially in high-risk areas like healthcare facilities. Portable UVGI units offer flexible deployment in hospitals and isolation rooms, often featuring fans to recirculate air through an internal UV chamber for targeted disinfection. These devices have demonstrated efficacy in reducing airborne microbial loads by up to 90% in recirculated air flows, making them valuable for point-of-care infection control where fixed installations are impractical. Validation of all UVGI air systems relies on airflow modeling to ensure dose uniformity, using or to predict UV exposure across varying velocities and mixing conditions, thereby optimizing lamp placement and output for consistent pathogen inactivation.

Water and Wastewater Treatment

Ultraviolet germicidal irradiation (UVGI) plays a critical role in drinking water treatment, particularly in large-scale municipal systems where open-channel reactors are employed to disinfect high volumes of water flowing through broad channels equipped with multiple UV lamps. These reactors facilitate efficient exposure of water to UV light without the need for pressure vessels, making them suitable for high-flow applications. For instance, the Catskill-Delaware Ultraviolet Disinfection Facility in utilizes 56 open-channel UV systems, each capable of delivering a 40 mJ/cm² dose to 40 million gallons per day, achieving a total capacity of approximately 2.2 billion gallons per day while ensuring effective inactivation of pathogens like and . Dose validation in drinking water treatment often relies on bioassays to confirm UV efficacy against resistant protozoa such as Cryptosporidium, with standard requirements specifying 10-20 mJ/cm² for at least 3-log inactivation (99.9% reduction) of oocysts, as determined through surrogate testing and direct infectivity assays. This range accounts for variations in water quality and reactor performance, ensuring compliance with regulatory standards for pathogen control. For wastewater treatment, UVGI systems typically employ medium-pressure lamps, which emit a polychromatic spectrum across 200-300 nm, providing broader germicidal action compared to low-pressure lamps and better penetration in turbid effluents with higher suspended solids. These lamps operate at elevated mercury vapor pressures, enabling higher intensities that mitigate the shielding effects of turbidity, a common challenge in secondary treated wastewater where total suspended solids can reduce UV transmittance to below 60%. Medium-pressure systems are thus preferred for reclaiming water or discharging to sensitive environments, achieving 4-6 log reductions in indicator bacteria like fecal coliforms at doses of 20-60 mJ/cm². Hybrid systems combining UVGI with chlorination address the lack of residual disinfectant in UV-treated water, using UV for primary inactivation followed by low-dose chlorine (0.2-0.5 mg/L) to maintain protection against recontamination during distribution or storage. This approach reduces overall chemical use while enhancing overall log reductions to 7-8 for viruses and bacteria in both drinking and wastewater contexts.

Surface, Food, and Specialized Uses

Ultraviolet germicidal irradiation (UVGI) is employed for direct surface disinfection in healthcare environments, utilizing portable devices such as robots and wands to target high-touch areas. During the 2020 COVID-19 pandemic, UVGI robots like the Xenex LightStrike system gained widespread adoption in hospitals, demonstrating the ability to inactivate on surfaces with exposures as short as 2 minutes, achieving reductions exceeding 4-log for viral loads. However, efficacy is limited by shadowing effects on irregular or occluded surfaces, where UV rays cannot reach, necessitating complementary manual cleaning. In food and beverage processing, UVGI serves as a non-thermal method for inline treatment of liquids like fruit juices, reducing pathogens without altering sensory qualities. The U.S. Food and Drug Administration (FDA) has approved UV irradiation under 21 CFR 179.39 for treating juices to achieve at least 5-log reductions of enteric pathogens such as Escherichia coli O157:H7 and Salmonella species, enabling preservation without heat pasteurization. For instance, commercial UV systems applied to apple cider and orange juice have demonstrated >5-log inactivation of E. coli at doses around 20 mJ/cm², maintaining nutritional content while extending microbial stability. Specialized UVGI applications encompass niche environments requiring precise microbial control, such as aquariums, cabinets, and outdoor s. In aquariums, UV sterilizers integrated into water circulation systems prevent algal blooms by inactivating free-floating planktonic algae like at doses of 10-30 mJ/cm², clarifying water without chemicals. cabinets often incorporate UV lamps for post-use surface , with 30-minute exposures recommended to achieve sufficient inactivation of residual contaminants on work surfaces, though this supplements rather than replaces chemical disinfection. For pond management, UV clarifiers in circulation loops control green water algae by delivering >30 mJ/cm² doses to flowing water, effectively reducing suspended microorganisms while preserving aquatic ecosystems.

UVGI Technologies

Mercury-Vapor and Low-Pressure Lamps

Mercury-vapor lamps have long served as the cornerstone of ultraviolet germicidal irradiation (UVGI) systems, particularly low-pressure variants that dominate traditional applications due to their reliable production of germicidal wavelengths. These lamps operate by exciting mercury vapor within a sealed tube using an , generating ultraviolet radiation primarily at 254 nm, which aligns with the peak absorption for DNA damage in microorganisms. Low-pressure mercury lamps emit approximately 85% of their ultraviolet output at 254 nm, making them highly selective for germicidal purposes, with the remainder including a minor peak at 185 nm that can produce if not suppressed. Their electrical-to-UV conversion efficiency typically ranges from 35% to 40%, outperforming many alternatives in energy use for monochromatic output. These lamps are constructed with a tube envelope to transmit UV-C , filled with mercury vapor and an such as , and equipped with electrodes; they require ballasts to regulate current and prevent instability in the discharge arc. Power output for standard low-pressure models is around 0.5 W/cm of arc length, while high-output versions can reach 1.5–10 W/cm. In contrast, medium- and high-pressure mercury lamps provide a broader spanning 200–300 , which enhances against UV-resistant organisms by targeting multiple bands beyond the 254 . These variants operate at higher vapor pressures, resulting in polychromatic output that includes significant energy in the 200–400 range overall, though focused germicidal contributions lie within 200–300 . They achieve higher power densities, typically 10–100 W/cm of , enabling compact designs but with reduced efficiency compared to low-pressure types due to greater heat and visible light production. Like their low-pressure counterparts, they use envelopes and require ballasts, but demand more robust cooling to manage elevated operating temperatures. A key operational characteristic of mercury-vapor lamps is their lifespan of 9,000–12,000 hours for low-pressure models, though medium-pressure variants last 4,000–8,000 hours due to intensified wear; output degrades gradually, often maintained at 80% after half the rated life. Startup requires a of 5–10 minutes to vaporize the mercury sufficiently for stable , particularly in preheat-start configurations common to low-pressure lamps, followed by a brief period to reach full intensity. These lamps offer cost-effectiveness through low initial and operational expenses, with widespread availability supporting their historical dominance in UVGI. However, they pose challenges related to mercury content, necessitating specialized disposal to mitigate environmental contamination, a concern amplified by the European Union's , which imposed restrictions on mercury in electrical equipment and led to phased exemptions for UV lamps, valid until at least 2027 and prompting gradual transitions in regulated markets.

Light-Emitting Diodes

Ultraviolet light-emitting diodes (UV LEDs) represent a solid-state alternative to traditional mercury-based lamps in ultraviolet germicidal irradiation (UVGI), offering compact and mercury-free sources for disinfection applications. These devices, primarily based on aluminum gallium nitride (AlGaN) semiconductors, emit in the UV-C range of 255-280 nm, with optimal germicidal performance peaking at around 265 nm due to strong absorption by microbial DNA. Tunability within this band allows customization for specific pathogens, enhancing efficacy without the broad-spectrum emissions of lamps. The design of UV-C LEDs leverages gallium nitride (GaN)-based heterostructures grown via metal-organic chemical vapor deposition, enabling compact chips with instantaneous activation and no warm-up time, unlike vapor lamps. Typical power output per chip ranges from 1 to 50 mW, sufficient for point-source applications, though scaling requires arrays for higher irradiance. External quantum efficiency (EQE) has progressed from below 5% in early commercial models to 10-20% by 2023 for wavelengths around 275 nm, with wall-plug efficiencies reaching 6% at 275 nm, driven by improvements in light extraction and defect reduction. Lifetimes exceed 10,000 hours at L70 (70% output retention), supporting reliable operation, though heat dissipation remains critical due to junction temperatures impacting performance. Compared to lamps, LEDs provide better dose uniformity in modular setups, reducing shadowing in irregular geometries. UV-C LEDs gained commercial viability around 2019 for niche disinfection uses, such as point-of-use , with costs dropping significantly to approximately $1,500 per watt by 2023—a reduction enabling broader adoption in portable devices like handheld wands. Key advantages include the absence of mercury, facilitating easier disposal and compliance with environmental regulations, alongside robustness to vibration and instant-on capability for intermittent use. However, limitations persist, including higher upfront costs (2-5 times that of lamps) and thermal management challenges that can degrade under prolonged operation. By , these trends have accelerated in and UVGI systems, with ongoing targeting 15% EQE to further close the gap with conventional sources.

Emerging Sources and Innovations

Far-ultraviolet C (far-UVC) lamps, particularly those utilizing krypton-chloride (KrCl) gas to emit at nm, represent a mercury-free advancement in UVGI sources designed for safe deployment in occupied spaces. These lamps produce narrowband radiation that inactivates airborne like aerosolized Staphylococcus aureus in room-sized environments without causing acute damage to or eyes when filtered to eliminate longer-wavelength emissions. Clinical evaluations, including a 2022 trial demonstrating over 99% pathogen reduction in a simulated indoor setting, confirm their efficacy for continuous air disinfection. Additionally, extended 36-month studies through 2024 have shown no significant ocular or dermal risks from prolonged exposure in full-room applications, supporting their use in healthcare and public venues. Filtered low-pressure mercury lamps emitting primarily at 184 nm, with ozone-producing wavelengths suppressed, offer another far-UVC variant tested in for occupied-space viability, achieving high microbial inactivation rates while minimizing human exposure hazards through optical filtering. Pulsed lamps provide high-intensity, broadband UV bursts (peaking in the UVC range alongside visible light) for rapid surface disinfection, inactivating human coronaviruses on healthcare surfaces in seconds via short-duration pulses that deliver cumulative doses exceeding 10 J/cm². These systems, often integrated into mobile or robotic platforms, have demonstrated log-6 reduction of bacteria like on food-contact surfaces, outperforming continuous sources in time-sensitive scenarios. Advancements in far-UVC light-emitting diodes (LEDs) are progressing toward scalability, with U.S. Department of Energy initiatives from targeting power conversion efficiencies above 15% by 2024 to enable widespread adoption, building on prototypes achieving around 1% for 222-230 nm emission. As of 2025, far-UVC LEDs have efficiencies of 1-5%, suitable for small-scale and point-source applications, while room-scale evaluations using lamps have demonstrated over 90% reductions in airborne microbial loads in occupied spaces at safe exposure levels below 3 mJ/cm² per hour. Emerging laser diodes at 222 nm enable precise, narrow-beam delivery for targeted disinfection, such as in confined medical equipment, with all-solid-state designs offering pulse capabilities up to 10 Hz for enhanced penetration without mercury. Filtered broadband sources, combining or outputs with dichroic filters to isolate far-UVC bands, further expand options for hybrid systems. Despite these innovations, challenges persist in cost, power scaling, and environmental integration. Far-UVC LEDs and excimer lamps remain 5-10 times more expensive than traditional mercury sources for equivalent output, limiting large-scale deployment, while power demands for whole-room systems often exceed 100 per unit, straining in retrofits. Mercury-free designs mitigate disposal hazards—reducing by over 90% compared to vapor lamps—but require advancements in materials like for sustained high-efficiency operation without degradation.

References

  1. [1]
    The History of Ultraviolet Germicidal Irradiation for Air Disinfection
    Ultraviolet germicidal irradiation (UVGI) is an established means of disinfection and can be used to prevent the spread of certain infectious diseases.
  2. [2]
    About Germicidal Ultraviolet (GUV) | Ventilation - CDC
    Oct 4, 2024 · GUV is the use of UV energy to kill viral, bacterial, and fungal organisms and is typically used in residential, commercial, educational, and healthcare ...Missing: authoritative | Show results with:authoritative
  3. [3]
    Germicidal Ultraviolet Air Treatment - Department of Energy
    A method of air and surface treatment that may be more effective and energy efficient to reduce airborne disease transmission than alternatives.Missing: authoritative | Show results with:authoritative
  4. [4]
  5. [5]
  6. [6]
    Repair of UV damage in bacteria - PubMed
    Mar 1, 2008 · These mechanisms include direct reversal of the damage by a photolyase (photoreactivation), removing of the damaged base by a DNA glycosylase ( ...
  7. [7]
    DNA Damage & Repair: Mechanisms for Maintaining DNA Integrity
    Bacteria and several other organisms also possess another mechanism to repair UV damage called photoreactivation. This method is often referred to as "light ...
  8. [8]
    Ultraviolet A Radiation - an overview | ScienceDirect Topics
    Ultraviolet light is classified as UV-A (320 to 400 nm), UV-B (280 to 320 nm) and UV-C (100 to 280 nm). UV-C is the most commonly used form of ultraviolet ...
  9. [9]
    [PDF] Action Of Uv-C Radiation for Biomolecules Inactivation, With ...
    Aug 4, 2023 · But the germicidal effectiveness of UVC peaks at about 260– 265 nm, which corresponds to the peak of UV absorption by bacterial DNA. The ...
  10. [10]
    (PDF) Ultraviolet Germicidal Irradiation Handbook - ResearchGate
    The germicidal effectiveness of UVC peaks at about 260–265 nm. This peak corresponds to the peak of UV absorption by bacterial DNA. The germicidal ...
  11. [11]
  12. [12]
    The impact of far-UVC radiation (200–230 nm) on pathogens, cells ...
    Feb 16, 2021 · Some researchers published reports that short-wavelength ultraviolet light in the spectral region of 200–230 nm (far-UVC) should inactivate pathogens without ...
  13. [13]
    Skin optical properties from 200 to 300 nm support far UV-C skin ...
    The results suggest that 233 nm far UV-C light emitting diodes (LEDs) could effectively inactivate microorganisms, while being safe and soft for the skin.
  14. [14]
    Critical discussion on the UV absorption properties of Earth's ...
    Below 200 nm the solar spectrum is strongly absorbed by molecular oxygen. The stratospheric ozone layer has strong absorption between 200 nm and 300 nm.
  15. [15]
    Why is UV radiation below 200nm strongly absorbed by Oxygen?
    Jul 15, 2023 · Ultraviolet radiation below 200 nm is considered to be Vacuum Ultraviolet, due to being strongly absorbed by atmospheric Oxygen.
  16. [16]
    Review Disinfection efficacy and safety of 222-nm ultraviolet C ...
    Some studies have suggested that 222 nm could be an alternative to 254 nm. Ha and colleagues' study indicated that the 222-nm UVC surface disinfecting ...
  17. [17]
    UV 254 nm is more efficient than UV 222 nm in inactivating SARS ...
    Wavelengths below 230 nm are generally highly effective for inactivation due to the absorbance of photons by both nucleic acids and proteins. Interestingly, ...
  18. [18]
    Suppression of photoreactivation of E. coli by excimer far-UV light ...
    May 15, 2024 · The maximum absorption of DNA is at 260 nm, while protein has absorption peaks at 280 and 230 nm (Beck et al., 2018; Beck et al., 2014), ...
  19. [19]
    Viability evaluation of layered cell sheets after ultraviolet light ... - NIH
    Most proteins show 10-fold more absorption coefficient at 222 nm than at 254 nm [16]. In case of spherical cells, nucleus and DNAs are covered with cytoplasm ...
  20. [20]
    UV FAQs - International Ultraviolet Association Inc
    ... fluence or UV dose). The "fluence" (UV dose) is obtained by multiplying the "fluence rate" (or irradiance) (units "mW/cm2") by the exposure time in seconds.Missing: equation | Show results with:equation
  21. [21]
    [PDF] ULTRAVIOLET DISINFECTION GUIDANCE MANUAL FOR THE ...
    A254 (UV Absorbance at 254 nm)– a measure of the amount of UV light that is absorbed by a substance at 254 nm. Action Spectra Correction Factor (CFas) – a ...
  22. [22]
    None
    Summary of each segment:
  23. [23]
    [PDF] Protocol for the determination of fluence (UV dose) using a low ...
    Bolton, J. R. and Linden, K. G. (2003), “Standardization of methods for fluence (UV dose) determination in bench-scale UV experiments”, J. Environ. Engr ...
  24. [24]
    Niels Ryberg Finsen – Facts - NobelPrize.org
    In 1895 he used concentrated beams of ultraviolet light to treat patients with lupus vulgaris with some success. For a time, light therapy was widespread, but ...Missing: 1880s- 1890s
  25. [25]
    Consuming light - Soaking Up the Rays - NCBI Bookshelf
    Koch demonstrated that sunlight could kill the tuberculosis bacilli in 1890: Robert Koch, Ueber bak-teriologische Forschung (Berlin:Hirschwald, 1890), cited ...
  26. [26]
    Walter Reed and the Scourge of Yellow Fever - UVA Today
    Nov 13, 2019 · UVA alumnus Walter Reed led the US Army Yellow Fever Commission in Cuba. The team proved that yellow fever was spread by mosquitoes.Missing: 1903 | Show results with:1903
  27. [27]
    Overview of Water Disinfection by UV Technology -A Review
    Jan 4, 2021 · 1910: Henri et al described the first commercial UV facility that used UV light to disinfect water in Marseilles, France. Overview of Water ...Missing: Hesketh | Show results with:Hesketh
  28. [28]
    Introduction - SpringerLink
    Jul 9, 2009 · The 1930 s saw the first applications of UV systems in hospitals to control infections (Wells and Wells 1936, Hart and Sanger 1939, Robertson et ...
  29. [29]
    [PDF] The Control of Acute Respiratory Illness by Ultra-Violet Lights
    Centers (Camp Sampson, New York, and Great Lakes), have indicated that ultra-violet irradiation of barracks resulted in a reduction of about 20 per cent in ...
  30. [30]
    [PDF] Maintenance of Upper-Room Germicidal Ultraviolet (GUV) Air ...
    FIXTURE AND LAMP CLEANING. Accumulation of dust on the reflective surfaces and lamps of upper-room GUV fixtures accounts for a significant decline in ...<|control11|><|separator|>
  31. [31]
    History of UV - VIQUA
    ... water), and in 1910, the first drinking water treatment system opened in Marseilles, France. The evolution of UV. 1930s: The first tubular lamps were ...Missing: Hesketh | Show results with:Hesketh
  32. [32]
    A Comprehensive Analysis of the UVC LEDs' Applications and ...
    Apr 13, 2022 · The UVC LEDs' decontamination effectiveness is as good as mercury vapor lamps. In some cases, LEDs even provide better results than conventional mercury vapor ...Missing: advancements | Show results with:advancements
  33. [33]
    Dose factors heavily into ultraviolet disinfection system design
    Jan 29, 2021 · UV-C LED sources for UVGI. Advances in semiconductor design have yielded LEDs based on aluminum gallium nitride (AlGaN) that emit UV-C light ...
  34. [34]
    Far UV-C radiation: An emerging tool for pandemic control
    Jun 10, 2022 · Far UV-C has been shown to be highly effective for inactivation of airborne pathogens; yet this same radiation has minimal potential to cause damage to human ...
  35. [35]
    Far-UVC (222 nm) efficiently inactivates an airborne pathogen in a ...
    Mar 23, 2022 · One potential solution is Krypton Chloride (KrCl) excimer lamps (often referred to as Far-UVC), which can efficiently inactivate pathogens, such ...Missing: alternative | Show results with:alternative
  36. [36]
    Far-UVC light (222 nm) efficiently and safely inactivates airborne ...
    Jun 24, 2020 · Germicidal ultraviolet light, typically at 254 nm, is effective in this context but, used directly, can be a health hazard to skin and eyes. By ...
  37. [37]
    Far-UVC light (222 nm) efficiently and safely inactivates airborne ...
    Jun 24, 2020 · Far-UVC light (207-222 nm) efficiently inactivates airborne coronaviruses, with 99.9% inactivation at low doses, and is expected to be similar ...
  38. [38]
    Use of Ultraviolet Germicidal Irradiation (UVGI) as a COVID ... - HDIAC
    In mid-May 2020, the Metropolitan Transportation Authority (MTA) in New York City (NYC), following the example of Chinese subways, began using UV lamps to ...
  39. [39]
    Performance of Ventilation, Filtration, and Upper-Room UVGI ... - MDPI
    To manage PM2.5 in the kitchen, HEPA and in-duct MERV13 filters were integrated into the ventilation system. Results showed that hybrid ventilation outperformed ...
  40. [40]
    UV Disinfection for COVID-19 - International Ultraviolet Association Inc
    "The IUVA has assembled leading experts from around the world to develop guidance on the effective use of UV technology, as a disinfection measure, to help ...Missing: 2022-2025 | Show results with:2022-2025
  41. [41]
    Far-UVC fix: The wavelength that harms viruses, not humans - SPIE
    Sep 1, 2025 · But as far-UVC LEDs edge closer to commercialization, their performance still has a way to go. For wall-plug efficiencies, commercial white ...
  42. [42]
    Efficacy of single pass UVC air treatment for the inactivation of ...
    Apr 26, 2022 · A portable, single-pass UVC air treatment device (air flow 1254 L/min) could effectively inactivate bioaerosols containing bacterial and viral indicator ...
  43. [43]
    Susceptibility of five strains of Cryptosporidium parvum oocysts to ...
    This study investigated the response of five strains of C. parvum to UV. A collimated beam apparatus was used to apply controlled doses of monochromatic (254 nm) ...Missing: D90 | Show results with:D90
  44. [44]
    Ultraviolet Light (UV-C): An Effective Tool to Mitigate Pandemic Spread
    Sep 8, 2021 · Summary of Ultraviolet dosage (254 nm) on Coronaviruses [11]. Microbe. D90 Dose (J/m2). Coronavirus. 6.6. Berne virus (Coronaviridae). 7.2.
  45. [45]
    Purified scrapie prions resist inactivation by UV irradiation - PubMed
    The development of effective purification protocols has permitted evaluation of the resistance of isolated scrapie prions to inactivation by UV irradiation ...
  46. [46]
    Inactivation of Viruses on Surfaces by Ultraviolet Germicidal Irradiation
    These results clearly indicate that dsRNA and dsDNA viruses are more resistant to UVGI than those of ssRNA and ssDNA viruses (UV doses for dsRNA and dsDNA was ...Test Viruses · Surface Test System · Results
  47. [47]
    [PDF] Inactivation of Pathogens in Air Using Ultraviolet Direct Irradiation ...
    Mar 1, 2022 · It is demonstrated herein that low-intensity UV radiation below exposure limits can achieve high levels of equivalent air changes per hour ( ...
  48. [48]
    Mathematical Modeling for Evaluating Inherent Parameters Affecting ...
    Mar 15, 2022 · The Chick-Watson model is the standard method for modeling the response of bacteria to UV decontamination (7). It explains the log-linear ...
  49. [49]
    Bacteria and RNA virus inactivation with a high-irradiance UV-A ...
    Sep 21, 2024 · UV-A at high values of irradiance (~ 0.46 W/cm2) can potentially induce ROS formation and direct photochemical damage of the pathogen nucleic ...
  50. [50]
    Determination of the characteristic inactivation fluence for SARS ...
    Jul 27, 2021 · In this work, we exposed SARS-CoV-2 to UV-C radiation ( = 254 nm) and we calculated—considering the influence of the optical absorption of the ...
  51. [51]
    Photoreactivation of Escherichia coli after Low- or Medium-Pressure ...
    The mechanisms by which UV light inactivates microorganisms are different at different wavelengths (14). The germicidal effect of short-wavelength UV light (UV- ...
  52. [52]
    Enhanced inactivation of E. coli by pulsed UV-LED irradiation during ...
    Feb 10, 2019 · This study demonstrated that pulsed UV-LED irradiation was more effective than continuous UV irradiation for E. coli inactivation.
  53. [53]
    Do Not Use Ultraviolet (UV) Wands That Give Off Unsafe ... - FDA
    Aug 17, 2023 · The type of eye injury associated with exposure to UV-C causes severe pain and a feeling of having sand in the eyes. FDA Actions. The FDA has ...
  54. [54]
    [PDF] FACT SHEET - Environmental Health And Safety
    UV exposure can cause lesions of the cornea and ultimately cause photokeratitis. Symptoms are described as a sensation of sand in the eye that may last for ...
  55. [55]
    Occupational Safety: UV Light Guidelines
    Aug 9, 2021 · The adverse health effects that may occur are erythema (sunburn) ... UV-B. 280-315nm. corneal injuries. cataracts of lens, photokeratitis.<|control11|><|separator|>
  56. [56]
    Toxic effects of ultraviolet radiation on the skin - ScienceDirect.com
    Chronic exposure to UV irradiation leads to photoaging, immunosuppression, and ultimately photocarcinogenesis. ... erythema, burns, and eventually skin ...
  57. [57]
    UV Radiation and the Skin - PMC - PubMed Central
    UV radiation (UV) is classified as a “complete carcinogen” because it is both a mutagen and a non-specific damaging agent and has properties of both a tumor ...
  58. [58]
    Exploratory clinical trial on the safety and bactericidal effect of 222 ...
    Aug 12, 2020 · This study aimed to investigate the safety of 222-nm UVC irradiation and to examine its skin sterilization effect in healthy volunteers.
  59. [59]
    Far-UVC light (222 nm) efficiently and safely inactivates airborne ...
    In fact there is a regulatory limit as to the amount of 222 nm light to which the public can be exposed, which is 23 mJ/cm2 per 8-hour exposure. Based on our ...
  60. [60]
    Sun-Safe Babies - The Skin Cancer Foundation
    Sep 13, 2016 · The skin of all infants (not just those who have lighter skin) is particularly vulnerable to sun damage.
  61. [61]
    Ultraviolet radiation - World Health Organization (WHO)
    Jun 21, 2022 · Fair-skinned people suffer more from sunburn and have a higher risk of skin cancer than dark-skinned people; however, darker-skinned people also ...Overview · Health Effects · Effects On The Skin
  62. [62]
    Protective goggles used during UV light therapy, England, 1930-1950
    High concentrations of UV light can damage the eyes, so green plastic goggles like these were worn by children undergoing ultraviolet light therapy or ...
  63. [63]
    Impact of UV-C on material degradation: a scoping literature review
    -Complete degradation after 16h UV-C (254 nm) exposure -Montmorillonite delays degradation rate -1H NMR/GPC results: --Polylactide chain scission → low MW ...
  64. [64]
    [PDF] Whitepaper: Ultraviolet (UV) Irradiation for Disinfection
    One recent paper (Heilingloh, et al., 2020) shows that a high dose of UVGI (1048 mJ/cm2) inactivated 99.9999% of SARS-CoV-2 in a highly concentrated viral ...
  65. [65]
    [PDF] DEFINING THE EFFECTIVENESS OF UV LAMPS - OSTI
    The 185-nm emission line for mercury is generated in low-pressure discharge lamps and would generate ozone in air. However, most low-pressure UVGI lamps ...
  66. [66]
    Unwanted Indoor Air Quality Effects from Using Ultraviolet C Lamps ...
    Ultraviolet germicidal irradiation (UVGI) is known to inactivate various viruses and bacteria, including SARS-CoV-2, and is widely applied especially in ...<|control11|><|separator|>
  67. [67]
    Indoor Air Quality Implications of Germicidal 222 nm Light
    Oct 12, 2023 · A promising new approach for GUV-based air cleaning is the use of KrCl excimer lamps, which emit at 222 nm (GUV222). (4) In contrast to 254 nm ...
  68. [68]
    [PDF] Energy Implications of Using Upper Room Germicidal Ultraviolet ...
    The reviewed studies found that the germicidal efficacy of UR-GUV equates to that of multiple air changes using pathogen-free air, and is more effective at ...
  69. [69]
    TLV Physical Agents Introduction - ACGIH
    This section presents Threshold Limit Values (TLVs) for occupational exposure to physical agents of acoustic, electromagnetic, ergonomic, mechanical, and ...Missing: permissible | Show results with:permissible
  70. [70]
    [PDF] FarUV KrCl Excimer lamps for disinfection
    ➢At 222 nm the ACGIH proposed eye TLV dose is 161 mJ/cm2. ✓The proposed ACGIH TLV dose for the skin is 478 mJ/cm2 (3 times larger). ✓The current ACGIH TLV ...<|separator|>
  71. [71]
    [PDF] on limits of exposure to ultraviolet radiation of wavelengths between
    Guidelines on limits of exposure to UV radiation ○ICNIRP. Page 5 ... upon the unprotected skin should not exceed 30 J m. ⫺2 effective spectrally ...
  72. [72]
  73. [73]
    Different biological effects of exposure to far-UVC (222 nm) and near ...
    Although nucleic acids can absorb and be modified by both UV 254 nm and UV 222 nm equally, compared to UV 254 nm, UV 222 nm is more intensely absorbed by ...
  74. [74]
    IES and IUVA release a new American National Standard for ...
    Mar 31, 2022 · The standard published today, ANSI/IES/IUVA LM-92-22 Approved Method: Optical and Electrical Measurement of Ultraviolet LEDs, details a repeatable method for ...<|separator|>
  75. [75]
    Institutional Tuberculosis Transmission. Controlled Trial of Upper ...
    Aug 15, 2015 · Upper room germicidal UV air disinfection with air mixing was highly effective in reducing tuberculosis transmission under hospital conditions.
  76. [76]
    (PDF) UV-C for HVAC Air and Surface Disinfection - ResearchGate
    Oct 13, 2020 · This article provides engineer-level guidance for the use of UV-C light to continuously reduce and even prevent the growth of dangerous microbes inHVAC systems.
  77. [77]
    Effectiveness of a UVC air disinfection system for the HVAC of an ICU
    Dec 18, 2021 · Tseng and Li assessed that UVGI doses of 2 to 5 mJ/cm2 were sufficient to inactivate single-stranded RNA viruses of the same type as SARS-CoV-2 ...Missing: 10-50 cm²
  78. [78]
    Upper-Room Ultraviolet Germicidal Irradiation (UVGI) - CDC Archive
    Apr 9, 2021 · Note: For airborne viral particles, upper-room UVGI systems provide air changes per hour that are similar to the introduction of clean air into ...
  79. [79]
    NYC will use powerful ultraviolet lamps to kill the coronavirus on ...
    May 4, 2020 · NYC will use powerful ultraviolet lamps to kill the coronavirus on subways and buses. The MTA wants to see if UVC rays can destroy the virus.
  80. [80]
    Applications of ultraviolet germicidal irradiation disinfection in health ...
    This review evaluates the applicability and relative contribution of ultraviolet germicidal irradiation (UVGI) to disinfection of air in health care facilities.
  81. [81]
    [PDF] Design and Optimization of UVGI Air Disinfection Systems - PSU-ETD
    The algorithms are based on models for 1) the intensity field of UVGI lamps, 2) the intensity field due to UVGI reflective enclosures, and 3) the kill rate of ...
  82. [82]
    Lagrangian modeling of inactivation of airborne microorganisms by ...
    Nov 21, 2020 · In this study, we applied the Lagrangian method to model the disinfection performance of in-duct UV lamps on three bacteria.Missing: uniformity | Show results with:uniformity
  83. [83]
    UV Treatment for Municipal Drinking Water - Trojan Technologies
    There are 56 custom-designed TrojanUV systems installed there and each of them is capable of delivering a 40 mJ/cm2 dose to 40 million gallons of water per day.Missing: open- reactors
  84. [84]
    [PDF] Wastewater Technology Fact Sheet Ultraviolet Disinfection
    Turbidity and total suspended solids (TSS) in the wastewater can render UV disinfection ineffective. UV disinfection with low-pressure lamps is not as effective.
  85. [85]
    UV Disinfection Systems for Wastewater Treatment: Emphasis on ...
    UV disinfection is cost-effective and easy to maintain for decentralized areas. However, to ensure its effectiveness, some parameters need to be considered.<|separator|>
  86. [86]
    2023 IUVA WC Posters - International Ultraviolet Association Inc
    Biofilm control with UV light: progress, research gaps ... UV light source for disinfection and photochemical oxidation in water and wastewater treatment.
  87. [87]
    Impact of combined chlorine and UV technology on the bacterial ...
    Hybrid disinfection has been studied with a sequential addition of UV followed by chlorine in drinking water and also chlorine followed by UV in wastewater ...
  88. [88]
    Xenex LightStrike Robot Destroys SARS-CoV-2 (Coronavirus) in 2 ...
    Apr 30, 2020 · Xenex LightStrike Robot Destroys SARS-CoV-2 (Coronavirus) in 2 Minutes; First & Only UV Disinfection Technology Proven to Deactivate COVID-19 ...
  89. [89]
    Verifiable Surface Disinfection Using Ultraviolet Light with a Mobile ...
    Mar 29, 2022 · By using robots in place of human health care workers in disinfection tasks, we can reduce the exposure of these workers to the virus and, as a ...
  90. [90]
    UV Disinfection Robots: A Review - PMC - PubMed Central
    Dec 9, 2022 · The UVC-LED system achieved 3-log inactivation of HCoV-OC43 at irradiation of 6–7 mJ/cm2 after 60 s. In addition, the UVC lamp was able to ...
  91. [91]
    A Review of the Efficacy of Ultraviolet C Irradiation for ... - NIH
    The authors reported that a 5 log reduction of E. coli and S. Typhimurium required UV-C treatment at a dose of 21 and 20 mJ/cm2, respectively [19]. In ...Missing: GRAS | Show results with:GRAS
  92. [92]
    Effects of UV-C Irradiation and Vacuum Sealing on the Shelf-Life of ...
    Feb 1, 2023 · It can extend the shelf life of various meat, poultry, and seafood products by 50–400%, which results in a reduction in losses, enabling the ...
  93. [93]
    UV Information - X Flo Systems
    » ALGAE, UV DOSE ; Chlorella vulgaris, 22 mJ/cm2 ; » BACTERIA ; Aeromonas salmonicida, 3.6 mJ/cm2 (log-3) ; Pseudomonas fluorescens (fin rot), 11 mJ/cm2 (log-3).Missing: prevention cm²
  94. [94]
    How much UV-light exposure time is required to disinfect a laminar ...
    Jul 3, 2013 · Typically 30 minutes to an hour are used for laminar flow hoods. However, it also depends on how old the bulb is. The older it is the weaker the irradiance.<|control11|><|separator|>
  95. [95]
    Algae treatment for UV disinfection
    To achieve this, the UV reactor must be correctly sized. The most important criterion for algae treatment is the dose applied: >30 mJ/cm². The next step is ...
  96. [96]
    Low Pressure Mercury Lamp: Top Products at LightSources
    They have a germicidal efficiency of 30% and can operate on lower current (between 180 and 425mA). The lamps come in both standard and high-output selections ...Missing: emission lifespan startup
  97. [97]
    [PDF] UV LAMP - Helios Quartz
    The low-pressure amalgam lamps produced by Helios are the best solution for ultraviolet disinfection and oxidation of water, air and surfaces because they.
  98. [98]
    UV medium-pressure lamps
    UV medium-pressure lamps feature a broad, pronounced line spectrum in the ultraviolet and visible range (200 nm to 600 nm) and a high power density of about ...Missing: UVGI | Show results with:UVGI
  99. [99]
  100. [100]
    RoHS Directive - Environment - European Commission
    It currently restricts the use of ten substances: lead, cadmium, mercury, hexavalent chromium, polybrominated biphenyls (PBB) and polybrominated diphenyl ethers ...RoHS Directive implementation · 2011/65 - EN - rohs 2 - EUR-Lex
  101. [101]
    A critical review of ultra-violet light emitting diodes as a one water ...
    Oct 30, 2024 · This literature review examines the progression of UV LED technologies from 2007 to 2023 using key features such as total optical power, price, and wall-plug ...Missing: IUVA | Show results with:IUVA
  102. [102]
    Highly efficient AlGaN-based deep-ultraviolet light-emitting diodes
    Aug 13, 2024 · We demonstrate the performance of 270-nm AlGaN-based DUV LEDs beyond the state-of-the-art by exploiting the innovative combination of bandgap engineering and ...
  103. [103]
    10.6% external quantum efficiency germicidal UV LEDs grown on ...
    Dec 4, 2023 · We report on the material challenges of the growth of highly conductive n-AlGaN in germicidal ultraviolet light emitting diodes (GUV LEDs)
  104. [104]
    Performance and reliability of state-of-the-art commercial UVC light ...
    State-of-the-art commercially available UVC LEDs have external quantum efficiencies below 4%. · Over 10,000 h L70 lifetime expected from many devices.
  105. [105]
    UV-LEDs: Next-Generation Ultraviolet Light Sources - IntechOpen
    Oct 28, 2025 · The EQE of UVC LEDs has been below 10% in the 265–280 nm band for a long time, mainly due to the presence of a large number of non-radiative ...
  106. [106]
    Ocular safety of 222‐nm far‐ultraviolet‐c full‐room germicidal ...
    Dec 10, 2024 · To evaluate the ocular safety of UV-C lamps for human exposure, we recently conducted a human study involving exposure to 222 nm far UV-C light.
  107. [107]
    Pulsed-Xenon Ultraviolet Light Highly Inactivates Human ... - NIH
    This study provides information on efficient and fast methods of disinfecting surfaces contaminated with different human coronaviruses (CoVs) in healthcare ...
  108. [108]
    Evaluation of pulsed xenon UV irradiation on inactivation of Listeria ...
    Jul 15, 2023 · UV-C disinfection is a safe and effective method for the inactivation of bacteria, viruses, spores, and other microorganisms. Pulsed xenon ...
  109. [109]
    [PDF] Solid-State Lighting Program: Germicidal Ultraviolet (GUV) R&D ...
    May 6, 2022 · 1) Improve UVC LED devices to achieve power conversion efficiency of > 15% by 2024 and > 30% by 2030, L70 lifetimes > 30,000 hours, and ...
  110. [110]
    [PDF] The effectiveness of Far-UVC UVGI-system vs an upper-room 253.7 ...
    Oct 7, 2025 · The newer Far-UVC technology, based on 222 nm radiation, irradiates the entire classroom space and offered a potentially safer and more ...
  111. [111]
    All-solid-state far-UVC pulse laser at 222 nm ... - Researching
    At present, the ultraviolet (UV) light sources used typically for disinfection mainly include low-pressure mercury lamps and UV light-emitting diodes (UV-LEDs).Missing: irradiation | Show results with:irradiation
  112. [112]
    Environmental life cycle assessment of UV-C LEDs vs. mercury ...
    This study is the first environmental comparison between a UV-C LED lamp (emitting at 265 nm) and mercury lamps employed in a lab-scale photoreactor for water ...Missing: UVGI | Show results with:UVGI