Photochemical & Photobiological Sciences (2023) 22:937–989
https://doi.org/10.1007/s43630-023-00371-y
PERSPECTIVES
Stratospheric ozone, UV radiation, and climate interactions
G. H. Bernhard1
· A. F. Bais2
· P. J. Aucamp3
· A. R. Klekociuk4
· J. B. Liley5
· R. L. McKenzie5
Received: 20 December 2022 / Accepted: 13 January 2023 / Published online: 21 April 2023
© The Author(s) 2023
Abstract
This assessment provides a comprehensive update of the effects of changes in stratospheric ozone and other factors (aerosols,
surface reflectivity, solar activity, and climate) on the intensity of ultraviolet (UV) radiation at the Earth’s surface. The assessment is performed in the context of the Montreal Protocol on Substances that Deplete the Ozone Layer and its Amendments and
Adjustments. Changes in UV radiation at low- and mid-latitudes (0–60°) during the last 25 years have generally been small (e.g.,
typically less than 4% per decade, increasing at some sites and decreasing at others) and were mostly driven by changes in cloud
cover and atmospheric aerosol content, caused partly by climate change and partly by measures to control tropospheric pollution.
Without the Montreal Protocol, erythemal (sunburning) UV irradiance at northern and southern latitudes of less than 50° would
have increased by 10–20% between 1996 and 2020. For southern latitudes exceeding 50°, the UV Index (UVI) would have surged by
between 25% (year-round at the southern tip of South America) and more than 100% (South Pole in spring). Variability of erythemal
irradiance in Antarctica was very large during the last four years. In spring 2019, erythemal UV radiation was at the minimum of
the historical (1991–2018) range at the South Pole, while near record-high values were observed in spring 2020, which were up
to 80% above the historical mean. In the Arctic, some of the highest erythemal irradiances on record were measured in March and
April 2020. For example in March 2020, the monthly average UVI over a site in the Canadian Arctic was up to 70% higher than the
historical (2005–2019) average, often exceeding this mean by three standard deviations. Under the presumption that all countries
will adhere to the Montreal Protocol in the future and that atmospheric aerosol concentrations remain constant, erythemal irradiance at mid-latitudes (30–60°) is projected to decrease between 2015 and 2090 by 2–5% in the north and by 4–6% in the south due
to recovering ozone. Changes projected for the tropics are ≤ 3%. However, in industrial regions that are currently affected by air
pollution, UV radiation will increase as measures to reduce air pollutants will gradually restore UV radiation intensities to those
of a cleaner atmosphere. Since most substances controlled by the Montreal Protocol are also greenhouse gases, the phase-out of
these substances may have avoided warming by 0.5–1.0 °C over mid-latitude regions of the continents, and by more than 1.0 °C
in the Arctic; however, the uncertainty of these calculations is large. We also assess the effects of changes in stratospheric ozone
on climate, focusing on the poleward shift of climate zones, and discuss the role of the small Antarctic ozone hole in 2019 on the
devastating “Black Summer” fires in Australia. Additional topics include the assessment of advances in measuring and modeling
of UV radiation; methods for determining personal UV exposure; the effect of solar radiation management (stratospheric aerosol
injections) on UV radiation relevant for plants; and possible revisions to the vitamin D action spectrum, which describes the wavelength dependence of the synthesis of previtamin D3 in human skin upon exposure to UV radiation.
* G. H. Bernhard
[email protected]
4
Antarctic Climate Program, Australian Antarctic Division,
Kingston, Australia
* A. F. Bais
[email protected]
5
National Institute of Water & Atmospheric Research, Lauder,
New Zealand
1
Biospherical Instruments Inc, San Diego, CA, USA
2
Laboratory of Atmospheric Physics, Department of Physics,
Aristotle University, Thessaloniki, Greece
3
Ptersa Environmental Consultants, Pretoria, South Africa
13
Vol.:(0123456789)
938
Photochemical & Photobiological Sciences (2023) 22:937–989
UVI Change 1996 – 2020
Graphical abstract
Change in UV Index during summer
100%
Observations of actual change
Range of expected UV changes without
the Montreal Protocol
50%
0%
-90
-60
-30
0
30
60
90
Latitude
Abbreviations
AAO
Antarctic Oscillation
AeroCom
Aerosol Comparisons between Observations and Models
AERONET
Aerosol Robotic Network
AO
Arctic Oscillation
AOD
Aerosol optical depth
asl
Above sea level
BUV
Backscatter Ultraviolet
CAMS
Copernicus Atmosphere Monitoring
Service
CAVA
Central American Volcanic Arc
CCM
Chemistry–climate model
CCMI
Chemistry–climate model initiative
CFC
Chlorofluorocarbon
CIE
Commission Internationale de l’
Éclairage (Eng.: International Commission on Illumination)
CL
Confidence level
CMIP6
Coupled Model Intercomparison Project
Phase 6
COVID-19
Coronavirus disease 2019
DSCOVR
Deep Space Climate Observatory
DTEC
((2Z,6Z)-2,6-bis(2-(2,6-diphenyl4H-thiopyran-4-ylidene)ethylidene)
cyclohexanone
EEAP
Environmental Effects Assessment Panel
EMAC
European Centre For Medium-Range
Weather Forecasts–Hamburg (ECHAM)/
Modular Earth Submodel System
(MESSy) Atmospheric Chemistry model
ENSO
El Niño-Southern Oscillation
EPIC
Earth Polychromatic Imaging Camera
EPP
Energetic particle precipitation
13
ERA
ETS
EUBREWNET
Geomip
GHG
GLENS
GOME
GPH
GWP
ICNIRP
HSRS
IPCC
NASA
MODIS
NDACC
NIWA
NOAA
NPP
NSF
ODS
OMI
OMPS
PAR
ECMWF (European Centre for MediumRange Weather Forecast) Re-analysis
Extraterrestrial (solar) spectrum
European Brewer Network
Geoengineering Model Intercomparison
Project
Greenhouse gas
Geoengineering Large Ensemble
Global Ozone Monitoring Experiment
Geopotential height
Global warming potential
International Commission on Non-Ionizing Radiation Protection
Hybrid Solar Reference Spectrum
Intergovernmental Panel on Climate
Change
National Aeronautics and Space Administration (of the United States)
Moderate Resolution Imaging
Spectroradiometer
Network for the Detection of Atmospheric Composition Change
National Institute of Water & Atmospheric Research (of New Zealand)
National Oceanic and Atmospheric
Administration (of the United States)
National Polar-orbiting Partnership
National Science Foundation (of the
United States)
Ozone-depleting substances
Ozone Monitoring Instrument
Ozone Mapping and Profiler Suite
Photosynthetically active radiation
(400–700 nm)
Photochemical & Photobiological Sciences (2023) 22:937–989
PM2.5
PPF
PSC
QASUME
QBO
RAF
RCP
RF
RSHU
SAI
SAM
SAP
SARS-CoV-2
SBUV
SED
SIM
SPE
SPF
SRM
SSA
SSP
SST
SSW
SURFRAD
SZA
TEMIS
TCO
TOMS
TROPOMI
TSI
TSIS
UNEP
USDA
UV
UV-A
UV-B
UV-C
UVI
VIS
VSLS
WMO
Particulate matter 2.5 (fine inhalable
particles, with diameters that are generally 2.5 µm or smaller)
Predictive protection factor
Polar stratospheric clouds
Quality Assurance of Spectral Ultraviolet Measurements in Europe
Quasi-biennial oscillation
Radiation Amplification Factor
Representative Concentration Pathways
Radiative forcing
Russian State Hydrometeorological
University
Stratospheric aerosol injection
Southern Annular Mode
Scientific Assessment Panel
Severe acute respiratory syndrome coronavirus 2
Solar Backscatter Ultraviolet
Radiometer
Standard erythemal dose
Spectral Irradiance Monitor
Solar proton events
Sun protection factor
Solar radiation management
Single scattering albedo
Shared Socioeconomic Pathways
Sea surface temperature
Sudden stratospheric warming
Surface Radiation Budget Network
Solar zenith angle
Tropospheric Emission Monitoring
Internet Service
Total column ozone
Total Ozone Mapping Spectrometer
Tropospheric Monitoring Instrument
Total solar irradiance
Total and Spectral Solar Irradiance
Sensor
United Nations Environment Programme
United States Department of Agriculture
Ultraviolet (100–400 nm)
Ultraviolet-A (315–400 nm)
Ultraviolet-B (280–315 nm)
Ultraviolet-C (100–280 nm)
Ultraviolet Index
Visible (radiation)
Very short-lived substances
World Meteorological Organization
939
1 Introduction
This Perspective is the first in a series of assessments1 prepared by members of the Environmental Effects Assessment
Panel (EEAP) of the Montreal Protocol under the United
Nations Environment Programme (UNEP). It focuses on the
effects of changes in the ozone layer on climate and ultraviolet (UV) radiation at the Earth’s surface, the interactions
between UV radiation and climate, and on the influence of
other geophysical parameters affecting UV radiation. This
assessment sets the stage for subsequent assessments in this
series that address the consequences of the interconnected
effects of stratospheric ozone depletion, UV radiation, and
climate change on human health [1] (including the COVID19 pandemic [2]), terrestrial [3] and aquatic [4] ecosystems,
the carbon cycle [3, 4], air quality [5], natural and synthetic
materials [6], and the fate of environmental plastic debris
[7]. These assessments focus on new scientific knowledge
that has accumulated since our last comprehensive assessment (Photochem. Photobiol. Sci., 2019, 18, 595–828) and
up to August 2022. Many of these effects are assessed in
terms of the benefits for life on Earth resulting from the
implementation of the Montreal Protocol on Substances
that Deplete the Ozone Layer [8] and its Amendments and
Adjustments (henceforth “the Montreal Protocol”). These
benefits were achieved by curbing depletion of stratospheric
ozone, thereby limiting increases of UV radiation, and mitigating climate change. Further topics include assessments of
observed trends in UV radiation, projections of UV radiation
into the future, and advances in the monitoring and modeling
of UV radiation.
2 State of the science in 2018
The previous comprehensive assessment of the EEAP [9],
which was based on the state of knowledge in 2018, concluded that the Montreal Protocol was highly beneficial for
protecting the stratospheric ozone layer and limiting the
rise of solar UV-B (280–315 nm) radiation at the Earth’s
surface. Therefore, increases in erythemal (sunburning) UV
radiation between the late 1970s (at the onset of anthropogenically induced stratospheric ozone depletion) and 2018
were negligible in the tropics, small (< 10%) at mid-latitudes
1
See the Photochem. Photobiol. Sci. themed issue entitled: Environmental effects of stratospheric ozone depletion, UV radiation,
and interactions with climate change: UNEP Environmental Effects
Assessment Panel, Quadrennial Assessment 2022 (https://doi.org/10.
1007/s43630-023-00374-9).
13
940
(30–60°), and large (> 50%) only in polar regions.2 Furthermore, the implementation of the Montreal Protocol3 prevented increases in UV-B radiation since the mid-1990s. As
a result, observed changes in UV radiation at mid-latitudes
during the last ~ 3 decades were mainly controlled by clouds
and aerosols instead of changes in stratospheric ozone. Statistically significant decreases in UV-B radiation consistent
with ozone recovery had not yet been detected at mid- and
low latitudes at the time of the previous assessment because
of the large variability in UV-B radiation caused by factors other than ozone. Conversely, continuing decreases in
clouds and aerosols (rather than changes in ozone) observed
since the mid-1990s led to positive trends of UV radiation
at several sites between 30° and 60° N. Several independent
satellite records indicated that changes in large-scale patterns of clouds occurred between the 1980s and 2000s with
consequences on UV radiation at the Earth’s surface.
In contrast to the tropics and mid-latitudes, variability of
UV-B radiation in Antarctica remained very large, with near
record-high erythemal UV radiation observed at the South
Pole in spring 2015 and well below average values in spring
2016. The Arctic remained vulnerable to large decreases in
total column ozone4 (TCO) and concomitant increases in
UV-B irradiance whenever meteorological conditions led to
a cold lower stratosphere in late winter and early spring. For
example, greatly reduced stratospheric ozone concentrations
during the second half of February 2016 led to increases of
erythemal UV radiation of up to 60% above the climatological average over northern Scandinavia and northern Siberia.
By preventing the further growth of the Antarctic ozone
hole, the Montreal Protocol also helped to reduce its effects
on atmospheric circulation, which include shifts of climate
zones in the Southern Hemisphere and associated changes
in weather patterns. For example, changes in tropospheric
circulation contributed to a decrease in summer temperatures
over south-east and south-central Australia, and inland areas
of the southern tip of Africa. Anomalously high TCO in the
spring were significantly correlated with hotter-than-normal
2
If not stated otherwise, the latitude ranges for both the Northern
and Southern Hemispheres are defined as: polar latitudes (80°–90°);
high-latitudes (60°–80°); mid-latitudes (30°–60°); low latitudes or
tropics (0°–30°).
3
The Montreal Protocol was adopted in 1987 and was implemented
in 1989 when it entered into force.
4
Total column ozone or TCO is the amount of ozone in a vertical
column extending from the Earth’s surface to the top of the atmosphere. TCO is reported in Dobson Units or DU. One DU corresponds
to a hypothetical layer of pure ozone with a thickness of 0.01 mm
that would ensue if all ozone molecules in the vertical column were
compressed to standard pressure (1013.25 hPa) and temperature
(273.15 K or 0 °C). One DU corresponds to 2.69 × 1016 molecules per
square centimeter of area at the base of this column. Averaged over
the Earth’s surface, the TCO is about 300 DU, which relates to a layer
of pure ozone that is three millimeters thick.
13
Photochemical & Photobiological Sciences (2023) 22:937–989
summers over large regions of the Southern Hemisphere and
vice versa.
With the predicted recovery of stratospheric ozone over
the next several decades, UV-B radiation was expected to
decrease at all latitudes outside the tropics, with the greatest decreases predicted over Antarctica. A projection of the
erythemal irradiance5 (quantified in terms of the UV Index6
or UVI) for the end of the twenty-first century (average
of 2085 − 2095) relative to the current decade (average of
2010 − 2020) suggested that ozone recovery will lead to a
decrease in the UVI by about 30% over Antarctica, and up
to 6% over mid-latitudes. These projections were uncertain
because future concentrations of stratospheric ozone will
depend not only on the decrease of ozone-depleting substances (ODSs) controlled by the Montreal Protocol but also
on the trajectory of concentrations of other greenhouse gases
such as carbon dioxide and methane, which will greatly
depend on policy decisions implemented in the coming decades. Changes in cloudiness were projected to result in small
(up to 4%) localized increases in UVI over the mid-latitudes
and tropics, and to decreases exceeding 10% in the Arctic.
Reductions in reflectivity due to melting of snow and sea
ice as well as shifting of the melting season were predicted
to decrease above-surface UVI by up to 10% in the Arctic
and by 2–3% around Antarctica. However, the increasingly
ice-free Arctic Ocean and reductions in snow cover would
lead to increases in UV radiation penetrating the water column and reaching land surfaces formerly covered by snow.
Decreases in concentrations of aerosols over urban areas of
the Northern Hemisphere were projected to increase the UVI
by typically 5–10% and by up to 30% over heavily industrialized regions (e.g., southern and eastern Asia) as measures
to control air pollution start to reduce contamination from
aerosols towards pre-industrial levels. The extent of these
changes was again determined to be greatly contingent on
policy decisions.
3 Current and future status of atmospheric
ozone
Changes in atmospheric ozone concentrations in general and
TCO in particular are regularly being assessed by the Scientific Assessment Panel (SAP) of the Montreal Protocol in
coordination with the World Meteorological Organization
(WMO) and UNEP. The information provided in this section
5
Irradiance is the radiant power (or radiant flux) received by a surface per unit area. “Radiant” indicates that the energy is received as
electromagnetic radiation, and the surface is assumed horizontal
unless otherwise specified.
6
The UV Index is calculated by weighting solar UV spectra with the
action spectrum of erythema [10] and multiplying the result with 40
m2/ W. See also Sect. 11.
Photochemical & Photobiological Sciences (2023) 22:937–989
941
(a)
(d)
(b)
(e)
(c)
(f)
Fig. 1 Time series of annual-mean TCO for the latitude bands a
35° N − 60° N, b 20° S − 20° N, and c 35° S − 60° S; and monthly
mean TCO for d March in the Arctic (60° N − 90° N), e September in the Antarctic (60° S − 90° S), and f October in the Antarctic
(60° S − 90° S). Colors indicate different ground- and satellite-based
datasets. These are identified in the legend of panel (b) and defined
as follows: WOUDC ground-based measurements from the World
Ozone and UV data center (https://woudc.org/); SBUV V8.7 NASA
(MOD): NASA Merged Ozone Data from the series of space-borne
Solar Backscatter Ultraviolet (SBUV) instruments; SBUV V8.6
NOAA (COH): the NOAA cohesive dataset from several satellite sensors; GOME/SCIA GSG: the merged dataset from the space-borne
Global Ozone Monitoring Experiment (GOME), the SCanning Imaging Absorption spectroMeter for Atmospheric CHartographY (SCIAMACHY), GOME-2A, and GOME-2B; and GOME/SCIA/OMI
GTO: the merged data set from GOME, SCIAMACHY, the Ozone
Monitoring Instrument (OMI), GOME-2A, GOME-2B, and TROPOspheric Monitoring Instrument (TROPOMI). The MLR (heavy
orange line) dataset is the median of the five datasets described above
and represents the input to the regression model applied by Weber
et al. [12]. Solid black lines indicate linear trends calculated with this
regression model before and after the peak in ODSs in 1996, respectively, and dotted lines indicate the two standard deviation (2σ) uncertainty of the estimated trends. Trend numbers are indicated for the pre
(1979–1995) and post (1996–2020) ODS peak period in the top part
of the plot. Numbers in parentheses are the 2σ trend uncertainty. The
dashed orange line shows the mean TCO from 1964 until 1980 from
the WOUDC data. Note that the scales of the ordinates are different
in the six panels. Adapted from Weber et al. [12]
is largely based on the SAP’s latest assessment [11] and
provides the background for our assessment of the various
effects resulting from changes in the ozone layer. We note
that trends in TCO assessed by the SAP and summarized
here refer to trends resulting mainly from human activities.
The effects of natural cycles and events that affect TCO have
been removed as part of the trend analysis. Such cycles and
events include the solar cycle; the quasi-biennial oscillation
(QBO; a pattern of alternating zonal winds in the tropical
stratosphere); the El Niño-Southern Oscillation (ENSO; a
pattern of alternating warm and cold sea surface temperatures of the tropical Pacific Ocean); the Arctic Oscillation
13
942
(AO) and the Antarctic Oscillation (AAO), which both
describe the back-and-forth shifting of atmospheric pressure
between the poles and the mid-latitudes; the Brewer–Dobson
circulation (a global-scale meridional circulation in the stratosphere); and aerosols from major volcanic eruptions [12].
3.1 Changes in total column ozone
outside the polar regions
Signs of the ozone layer’s recovery outside the polar regions
are now more robust compared to the SAP’s previous assessment [13] owing to updated trend models and additional
four years of data. For the first time, small but statistically
significant increases in TCO (of 0.4 ± 0.2% per decade) for
the period 1996–2020 are now evident for the latitude band
60° S–60° N [12]. However, this positive trend is mostly
driven by TCO changes in the Southern Hemisphere (Fig. 1).
In the tropics (20° S–20° N) and northern mid-latitudes
(35°–60° N), increases in TCO since 1996 have not been
observed with certainty (Fig. 1a, b), and statistically significant trends (of 0.7 ± 0.6% per decade) have only been
found for the southern mid-latitudes (35°–60° S) (Fig. 1c).
Even though the Montreal Protocol entered into force more
than 30 years ago, it was expected that the recovery of the
ozone layer at mid-latitudes would only now start to become
evident because the removal rate of ODSs controlled by the
Montreal Protocol from the stratosphere is three to four
times slower than the rate at which they were added [14].
Furthermore, year-to-year variability in TCO obscures the
attribution of trends to declining concentrations of ODSs.
Detecting significant increases in TCO outside Antarctica
therefore requires much more time than the detection of its
previous decline. In the upper stratosphere, however, the rate
of increase in the ozone concentrations is larger, ranging
between 1.5 and 2.2% per decade over the mid-latitudes of
both hemispheres, and between 1 and 1.5% per decade in
the tropics [11]. Since ozone column amounts in the upper
stratosphere (above 32 km) are relatively small (typically
less than 25% of the TCO at mid-latitudes), these increases
contribute only modestly to the growth of TCO. Over the
mid-latitudes, the present-day TCO (2018–2020 average) is
still below the average of the period 1964–1980 by ~ 4% in
the Northern Hemisphere and by ~ 5% in the Southern Hemisphere [11]. Reasons for these latitude-dependent changes in
TCO are discussed in SAP’s 2022 assessment [11].
3.2 Changes in total column ozone over Antarctica
Several studies have provided evidence that the Antarctic
ozone hole is starting to recover [15–21]. Signs of recovery
are strongest for the month of September, which is the key
month for chemical destruction of ozone. Both ground-based
and satellite data indicate a statistically significant positive
13
Photochemical & Photobiological Sciences (2023) 22:937–989
trend in TCO of 12% per decade in September since 2000
(Fig. 1e). These increases are consistent with the decrease in
the concentration of ODSs controlled by the Montreal Protocol [20]. However, there are still no significant trends for
October (Fig. 1f) or later months because TCO in late spring
is less sensitive to decreasing ODSs in the stratosphere compared to September. In a typical Antarctic winter, ozone is
almost completely destroyed in the lower stratosphere by
the end of September, which may explain why no recovery
has yet been observed in October over the polar cap [12].
In addition, year-to-year variability is also larger later in the
year [11].
Assuming continued adherence to the Montreal Protocol, concentrations of ODSs are projected to decline further,
eventually resulting in the disappearance of the annually
recurring ozone hole in the second half of the twenty-first
century [11]. Until that time, large year-to-year variations
in various ozone hole metrics are expected because of the
sensitivity of chemical ozone destruction to temperature in
the lower stratosphere in the presence of ODSs. Especially
during the last few years, the depth and size of the Antarctic
ozone hole have exhibited particularly large variability:
• In September and October 2019, the Antarctic ozone
hole was the smallest on record since the early 1980s
due to abnormally strong planetary wave7 activity originating in the subtropical Pacific Ocean east of Australia
and over the eastern South Pacific [22–24]. These waves
weakened the stratospheric polar vortex, which led to
a warming of the polar stratosphere, starting in midAugust [25]. The resulting above-normal temperature in
the lower stratosphere reduced the occurrence of polar
stratospheric clouds (PSCs), which provide the surfaces
for heterogeneous8 chemical reactions involving chlorine
that result in catalytic destruction of ozone. The volume
of PSCs dropped to almost zero by mid-September and
the chemical processes leading to ozone depletion were
therefore suppressed far earlier than usual. The average
TCO over the polar cap (60°–90° S) in September and
October 2019 was the highest over the last 40 years, and
the minimum TCO for September 2019 was the highest
since 1988. For the months of September, October, and
November, the polar cap average TCO was higher by
29%, 28%, and 26%, respectively, compared to the mean
of the 2008–2018 period [26].
• In contrast, the Antarctic ozone holes in spring 2020 and
2021 were amongst the largest and longest-lived in the
observational record [27, 28]. These long-lasting ozone
7
Large-scale perturbations in atmospheric circulation, typically
manifesting as meandering of the jet stream.
8
Heterogeneous chemical reactions are chemical reactions between
substances of different phases, e.g., gaseous, liquid, solid.
Photochemical & Photobiological Sciences (2023) 22:937–989
(a)
943
(b)
Fig. 2 a Time series of TCO at King George Island (62° S), averaged
from 1 September to 15 October (red line) and from 16 October to 30
November (blue line). b Evolution of the ozone hole area averaged
from 1 September to 15 October (red line) and from 16 October to
30 November (blue line). Bold lines indicate 11-year centered moving
averages calculated from annual data. Adapted from Cordero et al.
[34]
holes, extending to times when snow has melted, may
have had impacts on Antarctic organisms [29]. Yook
et al. [28] provided evidence that injection of smoke originating from the Australian “Black Summer” wildfires of
early 2020 (Sect 5.1.2) may have contributed to the large
ozone hole of 2020, while aerosols from the eruption
of La Soufrière (13° N) on Saint Vincent in April 2021
may have played a role in the large ozone hole of 2021.
(Aerosols injected into the tropical stratosphere disperse
rapidly to high latitudes [30].) Furthermore, the lack of
planetary waves during both years resulted in a cold and
stable stratospheric vortex over Antarctica, which created conditions favorable for persistent ozone depletion
[11, 20, 31]. Additionally, loss of ozone in early spring
2020 enhanced the strength and persistence of the vortex
later in that year [32]. Even though large ozone holes
will likely continue to occur in the future, either through
dynamical variability alone, or exacerbated by large volcanic eruptions or major inputs of smoke into the stratosphere, the recovery of the ozone hole is expected to
continue [27].
(Sect. 6) lead to additional variability, hampering detection
of recovery further.
Using observations from satellites between 1978 and
2020, a recent study [34] compared annual averages of the
depth and area of the Antarctic ozone hole for early spring
(1 September–15 October) and late spring (16 October–30
November). This analysis is of high relevance for assessing
trends in UV radiation over Antarctica because UV radiation
is generally much greater later in spring when the Sun is
higher in the sky even though TCO is typically much lower
earlier in spring. Figure 2a shows TCO averaged from 1
September to 15 October (red line) and from 16 October
to 30 November (blue line) at King George Island (62° S),
located near the northern tip of the Antarctic Peninsula.
For the earlier period, the 11-year moving average of TCO
was lowest around the year 2000, when the concentration
of ozone-depleting chlorine and bromine compounds in the
stratosphere was close to its maximum, and average TCO
appears to be increasing since this time. The observation at
this station is consistent with the positive trend in Antarctic
TCO for September shown in Fig. 1e. Conversely, and consistent with Fig. 1f, there is no clear indication that TCO is
also recovering in the later period. Similarly, the size of the
ozone hole—quantified as the area with TCO below 220
Dobson Units (DU)—appears to be decreasing faster in early
spring (Fig. 2b).
The large year-to-year variability in the TCO observed
thus far resulted in large year-to-year variations in UV radiation in Antarctica (Sect. 7.1.1). For example, the UVIs measured at the South Pole in 2019 were some of the lowest since
the start of measurements in 1991, while those in 2020 set
new record highs. The recovery of the Antarctic ozone hole
is generally more difficult to detect with UV-B radiation than
ozone data because signs of recovery are most pronounced
in September [15, 33] when the UVI in Antarctica is still
very low. Factors other than ozone that affect UV radiation
3.3 Changes in total column ozone over the Arctic
While there is still no clear evidence of ozone recovery in
the Arctic, it is expected that signs of recovery would first be
detected in March because chemical ozone loss in the Arctic
is typically largest in this month [35].
13
944
Figure 1d indicates that TCO in March averaged over the
northern polar cap (63°–90° N) is indeed increasing by 2%
per decade, but this small positive trend is not statistically
significant because of the large interannual dynamical variability observed for this latitude belt [11].
Sporadic ozone depletion events continue to occur in
the Arctic. An exceptionally large episode of stratospheric
ozone depletion was observed in late winter and early
spring (February–April) of 2020 [36], exceeding in severity the previously reported event of 2011 [37]. The TCO
averaged over 63°–90° N for this 3-month period was 340
Dobson Units (DU), which is 100 DU below the mean of
the period 1979–2019 and the lowest since the start of satellite measurements in 1979. These low values of TCO in
2020 were partially caused by a strong and long-lived polar
vortex, which provided ideal conditions for chemical ozone
destruction to take place. Temperatures low enough to form
PSCs within the vortex developed early in the season, and
on average enclosed about a third of the vortex volume [35,
36, 38–41]. Furthermore, the strong vortex also inhibited
replenishment of Arctic ozone from lower latitudes [11].
These conditions are unique in the ~ 40 years of measurements, making 2020 the year with the largest loss of Arctic ozone on record. The large ozone hole observed over
Antarctica six months later is a coincidence and cannot be
attributed to a known common cause.
The unprecedented depletion of Arctic ozone in winter/
spring of 2019/2020 contrasts with the conditions in the
boreal winters of 2018/2019 and 2020/2021. In both winters, major stratospheric warmings occurred in January
[42–44], which limited overall ozone loss. As a result, the
minimum TCO in March 2019 (defined as the minimum
of the daily mean TCO within an area that encloses the
Arctic polar vortex and is surrounded by the 63° N contour
of “equivalent latitude” [45]) was the highest since 1988
[46], and the minimum TCO in March 2021 was identical
to its average value since the start of satellite observations
in 1979 [47]. Such large year-to-year variations in Arctic
ozone depletion, which are driven by differences in meteorological conditions, are expected to continue for as long
as concentrations of ODSs remain elevated [11, 41, 48]. Furthermore, winters with a warm stratosphere (and little ozone
depletion) will likely randomly alternate with winters with a
cold stratosphere (and large ozone depletion). A recent study
[49] provides evidence that years with a cold stratospheric
Arctic vortex are getting colder. Reduced stratospheric
temperatures will likely result in more PSC formation and
lead to more chemical ozone loss via catalytic processes.
As a consequence, ozone-depletion events as large or even
larger than the one observed in 2020 [e.g., 36] will likely
re-occur throughout the twenty-first century until concentrations of ODSs have substantially decreased. The magnitude
of stratospheric cooling in the future will critically depend
13
Photochemical & Photobiological Sciences (2023) 22:937–989
on the development of greenhouse gas (GHG) concentrations and on variability in the amount of water (H2O) vapor
in the stratosphere [11, 49]. Under the scenario with the
highest concentration of GHGs and H2O, sporadic springtime increases in UV radiation in the Arctic could be somewhat larger at the end of the twenty-first century than those
observed in 2020 [49].
3.4 Effects of greenhouse gases on stratospheric
ozone
This section briefly discusses the effects of changes in the
atmospheric concentration of GHGs that are responsible for
global warming but are also relevant to stratospheric ozone
changes. The SAP’s latest report [11] discusses these processes in more detail. Increases in GHGs affect ozone depletion in several key ways [50]. First, radiative cooling of the
polar stratosphere (promoted by GHGs during winter months)
enhances the formation of PSCs. These clouds provide the
surfaces for heterogeneous chemical reactions that lead to the
destruction of ozone, thereby decreasing ozone concentrations. Second, cooling of the upper stratosphere at extrapolar
latitudes reduces the rates of gas-phase chemical reactions that
lead to ozone loss, thereby increasing ozone concentrations in
the upper stratosphere. Third, changes in the concentrations
of nitrous oxide (N2O) and methane (CH4), which are both
GHGs, also affect ozone concentrations chemically because
both gases are also key sources of reactive species in catalytic
cycles (the NOx and HOx cycles, respectively) that destroy
ozone. The NOx cycle dominates in the middle stratosphere
(approximately 25–35 km) while the HOx cycle is mostly contributing in the lower stratosphere. Fourth, increases in GHG
concentrations are expected to strengthen the Brewer–Dobson
circulation, which describes the redistribution of ozone from
tropical to extratropical regions [51]. Fifth, global warming
induced by increases in GHGs increases the flux of “very
short-lived substances” (VSLS) into the stratosphere as further explained in the following.
VSLS are ozone-depleting halogen-containing substances
with a lifetime of less than six months that are mostly produced
by natural processes, for example, by macroalgae (seaweed)
and phytoplankton. About 25% of bromine entering the stratosphere in 2016 was from VSLS [13], with the majority originating from oceanic sources. While stratospheric bromine is a
relatively minor constituent by volume, it is an important contributor to ozone depletion. Per atom, bromine is about 60–75
times (depending on the concentration of GHGs) more effective
in destroying ozone than is chlorine [52]. A recent modeling
study [53] examined the effect of climate change on changes
in bromine from oceanic sources. The study assumed the
Photochemical & Photobiological Sciences (2023) 22:937–989
Representative Concentration Pathway9 RCP 6.0 GHG scenario
and concluded that the flux of brominated VSLS compounds
from the ocean to the atmosphere will increase by about 10%
over the twenty-first century for all latitudes with the exception
of the Arctic. The increase will be even greater over the Arctic
because of the projected decrease in sea ice, which is currently
hindering the escape of brominated compounds from the ocean.
By the end of the twenty-first century, almost the entire polar
ocean will likely be exposed in August and September and sea
ice will no longer curtail ocean–atmosphere fluxes of brominated compounds. This study is one example of an indirect
effect of climate change on the concentration of substances that
promote stratospheric ozone depletion.
3.5 Estimates of total column ozone
during the twenty‑first century
Projections of TCO into the future are available from
chemistry-climate models (CCMs), which were run for different future emissions scenarios as part of a coordinated,
multi-model activity where all models follow the same
protocols to perform a comparable set of simulations [11,
54]. Uncertainties associated with these projections arise
mainly from the assumed future trajectories of emissions
of GHGs and pollutants. The models were run in the framework of CMIP610 simulations and follow a new set of future
emissions scenarios, the Shared Socioeconomic Pathways
(SSP11) [55], which assume compliance with the Montreal
Protocol and its Amendments. The ozone projections for the
different SSPs are therefore based on the same evolution of
9
Representative Concentration Pathways are greenhouse gas concentration (not emission) trajectories adopted by the Intergovernmental
Panel on Climate Change (IPCC) for its fifth Assessment Report. The
pathways are used for climate modeling and research. They describe
four climate futures, which differ in the amount of greenhouse gases
that are emitted in years to come. The four RCPs, RCP 2.6, RCP
4.5, RCP 6, and RCP 8.5, are named after a possible range of radiative forcing values in the year 2100 relative to pre-industrial values
(+ 2.6, + 4.5, + 6.0, and + 8.5 W m−2, respectively).
10
Coupled Model Intercomparison Project Phase 6.
11
Shared socio-economic pathway (SSP) scenarios describe a range
of plausible trends in the evolution of society over the twenty-first
century and were adopted by the Intergovernmental Panel on Climate
Change (IPCC) for its sixth Assessment Report. The pathways are
used for climate modeling and research, as different socio-economic
developments and political environments will lead to different GHG
emissions and concentrations. They describe five climate futures
(SSP1–SSP5) that are combined with assumed amounts of greenhouse gases that are emitted in years to come. The CMIP6 simulations are based on seven SSPs (SSP1-1.9, SSP1-2.6, SSP2-4.5, SSP37.0, SSP4-3.4, SSP4-6.0, and SSP5-8.5), which are named after a
possible range of radiative forcing (see footnote 14) values in the year
2100 relative to pre-industrial values (1.9, 2.6, 4.5, 7.0, 3.4, 6.0, and
8.5 W m−2, respectively), and have some equivalence to the “Representative Concentration Pathways” or RCPs used in IPCC’s fifth
Assessment Report.
945
controlled ODSs and depend only on the evolution of GHGs
and other pollutants.
The new simulations for the evolution of TCO towards
the year 2100 support conclusions similar to those presented
in a previous assessment of the SAP [13]. Figure 3 depicts
the evolution of the annual-mean TCO averaged over different latitude bands for the period 1950–2100. The projections
are based on a set of CMIP6 CCMs, which were run for the
historical period 1950–2015 as well as for different scenarios
for the future period 2015–2100. Year-to-year variability in
these simulations is the result of internal variability (sometimes called “weather noise” [13]).
In summary, for scenarios with stabilizing or slightly
decreasing concentrations of GHGs (SSP2-4.5, SSP4-3.4,
and SSP4-6.0), the near-global mean (60° S–60° N) TCO is
projected to return to historic levels (year 1980) by the middle of the twenty-first century (around year 2040) and remain
at those levels until 2100. For scenarios with continued GHG
increases (SSP3-7.0 and SSP5-8.5), the TCO is projected to
return to 1980 levels sooner and significantly exceed historic
levels throughout the latter half of the twenty-first century.
This overshoot, which has also been termed “super-recovery”, results from the fact that increases in GHGs cool the
upper stratosphere. This cooling reduces the rates of gasphase chemical reactions that destroy ozone, and as a result,
ozone concentrations increase. In contrast, and despite the
assumption that halogenated ODSs will continue to decline
throughout this century, TCO is not projected to return to
historic levels by 2100 for scenarios with small GHG emissions (SSP1-1.9 and SSP1-2.6) and is projected to decrease
in the tropics [11]. The consequences of these changes in
TCO on UV radiation at the Earth’s surface, and its dependence on the GHG scenario, are discussed in Sect. 8.
4 Benefits of the Montreal protocol
Benefits of the Montreal Protocol can be both direct (curbing stratospheric ozone depletion and limiting increases of
UV radiation) and indirect (effects on climate). This section
provides new information on both benefits.
4.1 Direct effects of the Montreal protocol
on stratospheric ozone depletion and UV
radiation
The phase-out of ODSs mandated by the Montreal Protocol
has already limited increases in UV radiation at the Earth’s
surface. To demonstrate this beneficial effect, McKenzie
et al. [56] compared seasonal means of the daily maximum
UVI measured at the Earth’s surface with UVI data derived
from results of two CCMs that assumed either the “World
Avoided” scenario, where emissions of ODSs would have
13
946
Photochemical & Photobiological Sciences (2023) 22:937–989
Fig. 3 Regional average CMIP6 multi-model annual-mean TCO for
the historical period (1950–2015) (black line), and the future (2015–
2100) based on seven SSP scenarios (colored lines). The six panels
show results for different latitudinal bands, indicated in the top left
of each panel. The number of models participating in each simulation is shown in parentheses in the legend. The light gray envelope
indicates the model spread for the historical simulations (calculated
as the standard error). Total ozone columns for the 1960 and 1980
annual means are given by the solid and dashed horizontal gray lines,
respectively. Note that the scales of the ordinates are different in the
six panels. Reprinted from Keeble et al. [54]
continued without regulation, or the “World Expected”
scenario, where ODSs are curbed in compliance with the
Montreal Protocol and its Amendments. The ground-based
measurements were made at 17 mostly clean-air sites (latitude range 73° N–90° S) by state-of-the-art spectroradiometers. Trends in the UVI over 1996–2018 derived from measurements at sites with sufficiently long data records were
found to be either small (< ± 10% per decade at Antarctic
sites) or not significantly different from zero. These estimates matched calculations following the World Expected
scenario within the limits of the measurement uncertainty. In
contrast, without the Montreal Protocol, the UVI at northern
and southern latitudes of less than 50° would have increased
by 10–20% between the early 1990s and 2018. For southern
latitudes exceeding 50°, UVI values would have surged by
between 25% (year-round at the southern tip of South America) and more than 100% (South Pole in spring and summer).
Figure 4 shows an update of the work by McKenzie et al.
[56] including also UVI measurements from 2019 and 2020,
and focusing on sites with at least 15 years of observations
between 1996 and 2020. With the exception of Thessaloniki
(41° N), changes in the UVI over this time period have been
smaller than ± 11% at all sites for both summer (Fig. 4a)
and spring (Fig. 4b), and smaller than the “World Avoided”
scenarios projected by the two CCMs (GEOSCCM12 [57]
and NIWA-UKCA13 [58]), confirming that the Montreal
Protocol has prevented large increases in UV radiation, in
particular at southern latitudes higher than 60°. For example,
without the Montreal Protocol (blue lines in Fig. 4), the UVI
at the South Pole would by now have more than doubled
in spring, while the ground-based measurements indicate
13
12
Goddard Earth Observing System Chemistry–Climate Model.
REF-C2 simulation of the NIWA-UKCA model (Implementation
of the United Kingdom Chemistry and Aerosols (UKCA) model by
New Zealand’s National Institute of Water & Atmospheric Research
(NIWA)) [58] with exponentially increasing concentrations of ODSs
at 3% per year added from 1974 onwards.
13
UVI Change 1996 – 2020
150%
UVI Change 1996 – 2020
Photochemical & Photobiological Sciences (2023) 22:937–989
150%
Summer
947
(a)
Observations
Modelled UVI change
without Montreal Protocol:
GEOSCCM model
NIWA-UKCA model
100%
50%
0%
-50%
-90
-60
-30
0
30
60
Spring
90
(b)
100%
50%
0%
-50%
-90
-60
-30
0
30
60
90
Latitude
Fig. 4 Comparison of relative changes in the UVI between 1996 and
2020 for a summer and b spring, derived from observations at nine
ground stations (black symbols) and calculated from results of two
chemistry-climate models (blue lines). Both climate models assume
the “World Avoided” scenario where emissions of ozone-depleting
substances are not controlled by the Montreal Protocol. Blue shading indicates the range of these model projections. Ground stations
include South Pole (90° S), Arrival Heights (78° S), Palmer Station
(65° S), Lauder (45° S), Alice Springs (24° S), Mauna Loa (20° N),
Boulder (40° N), Thessaloniki (41° N), and Barrow (71° N). Ground
stations with a near-complete data record for 1996–2020 are indicated
by solid symbols. Sites with less than 24 years of data are shown with
open symbols. Error bars indicate the 95% confidence interval of the
regression model. Updated from McKenzie et al. [56]
a decrease of 10 ± 34% (± 2 standard deviations). Projected
changes for high latitudes in the Northern Hemisphere are
generally smaller because ozone depletion over the Arctic
is less severe than that over the Antarctic (Sec 3.2 and 3.3).
The relatively large increases in the measured UVI at Thessaloniki (16% for spring and 8% for summer) are mostly
caused by reductions in atmospheric aerosols at this urban
site resulting from air pollution control measures (Sect. 6.1)
and are not the result of decreases in ozone.
halocarbons has greatly increased during the last century.
For example, over the second half of the twentieth century,
the combined direct radiative effect of all ODSs was the
second largest contributor to global warming after CO2,
with approximately one third of the radiative forcing14 (RF)
of CO2 [59]. The climate effects of ODSs were already
anticipated during the establishment of the Montreal Protocol [60], and their impact on climate has been continuously revised since the ratification of the Montreal Protocol
[13, 61, 62]. Work on assessing the contribution of ODSs
to global warming has continued during the last four years;
however, the net effect of ODSs on global temperatures is
still highly uncertain [Chapters 6 and 7 of 63] because some
of the warming that ODSs induce is offset by their effect on
stratospheric ozone. Specifically, since ozone is also a GHG,
depletion of ozone caused by ODSs has a cooling effect, but
the magnitude of this effect (hereinafter termed “indirect
forcing from ozone depletion”) is uncertain. On one hand,
two single-model studies have reported a very large cancelation of the direct forcing by ODSs by the indirect forcing
from ozone depletion of up to 80% [64, 65], and two multimodel studies using an “emergent constraint approach”15
based on CMIP6 models came to a similar conclusion [66,
67]. On the other hand, additional studies, which were part
of several model intercomparison projects, concluded that
the climatic effect from ODS-induced ozone depletion is
either small or negligible [68–72]. According to Chiodo
and Polvani [72], the four studies that have calculated a
large effect on climate from ozone depletion have weaknesses (e.g., one study was based on a short time period, one
study had a large ozone bias, and the remaining two studies assumed unrealistically strong ozone depletion), while
the other studies that indicate a small indirect forcing from
ozone depletion are more reliable because they are consistent with multi-model means of the CMIP5 and CMIP6 models, as summarized by Checa‐Garcia et al. [68]. However at
this time, results from the two groups of studies cannot be
reconciled. Because of these discrepancies, the latest (6th)
report of the Intergovernmental Panel on Climate Change
(IPCC) [specifically, Chapter 7 of 63] does not attempt to
quantify the indirect forcing from ozone depletion, in contrast to previous IPCC reports [e.g., 73].
In the following, we summarize the results of recent studies that evaluate the amount of global warming that has been
14
4.2 Indirect effects of the Montreal protocol
on climate
Most ODSs controlled by the Montreal Protocol are also
potent GHGs with Global Warming Potentials (GWPs) that
are substantially larger than those of carbon dioxide (CO2)
on a molecule-by-molecule basis. The climate forcing of
Radiative forcing quantifies the change in Earth’s energy balance
(in W m−2) between incoming short-wave solar radiation and outgoing long-wave (thermal) IR radiation, either at the tropopause or at
the top of atmosphere. If radiative forcing is positive at the tropopause, the temperature of the troposphere will increase.
15
Emergent constraints are physically explainable empirical relationships between characteristics of the current climate and long-term
climate prediction that emerge in large ensembles of climate model
simulations.
13
948
avoided due to the Montreal Protocol’s control of ODSs. All
studies implicitly calculate the indirect forcing from ozone
depletion and take this forcing into account when computing
the net effect of the Montreal Protocol on surface temperatures. However, because of the uncertainty in calculating
this feedback, the resulting effect on temperature is also
uncertain. Still, taken together, these new studies further
demonstrate the effectiveness of the Montreal Protocol in
limiting temperature rise at the Earth’s surface.
G oya l et a l . [ 7 4 ] u s e d a c o u p l e d a t m o s phere–ocean–land–sea ice model to re-evaluate the Montreal Protocol’s effect on global warming from the control
of ODSs. The study considered ODSs that have contributed substantially to stratospheric chlorine concentrations,
namely the chlorofluorocarbons (CFCs) CFC-11 and CFC12, as well as the CFC substitutes HCFC-22, HFC-125 and
HFC-134a. Increases in GHG concentrations (including
the concentrations of these ODSs) were described in this
model by RCP 8.5, which leads to the strongest warming at
the surface of the Earth. The study determined that, as of
2019, the Montreal Protocol has avoided warming between
0.5 to 1.0 °C over mid-latitude regions of Africa, North
America, and Eurasia and as much as 1.1 °C warming in
the Arctic. In addition to quantifying the benefits from the
Montreal Protocol that have already been realized, Goyal
et al. [74] also assessed the Montreal Protocol’s effect on
the future climate for the RCP 8.5 scenario. Projected temperature increases that are likely to be averted by 2050
are in the order of 1.5 °C–2 °C over most extrapolar land
areas, and between 3 °C and 4 °C over the Arctic. Averaged over the globe (including the oceans), about 1 °C
warming would be avoided by 2050, which corresponds
to about 25% mitigation of global warming expected from
all GHGs.
A separate study [59] found that, over the period
1955–2005, ODSs were responsible for about one third of
warming globally and about half of the warming in the Arctic. Since changes in Arctic temperatures have a direct effect
on sea ice loss, Polvani et al. [59] concluded that ODSs contributed half of the forced Arctic sea ice loss in the latter
half of the twentieth century. These results were recently
confirmed [75], showing that Arctic warming and sea ice
loss from ODSs are slightly more than half (52–59%) of
those from CO2.
More recently, Chiodo and Polvani [72] calculated that
stratospheric ozone depletion from ODSs only cancels about
25% of the RF from ODSs, in agreement with recent studies
[e.g., 68]. The net RF of ODS is 0.24 W/m2 accordingly,
which amounts to nearly one third of the RF of CO2 over the
period 1955–2005, emphasizing the large RF effect of ODSs
on tropospheric temperatures.
In summary, recent model calculations demonstrate a
large effect of the Montreal Protocol in limiting global
13
Photochemical & Photobiological Sciences (2023) 22:937–989
warming, but these results are subject to large uncertainties because the cooling effect resulting from ODS-induced
ozone depletion is quantitatively not well reproduced by
CCMs. The influence of ODSs on climate is an area of
active research and it is expected that refinements to chemistry-climate models will further reduce uncertainties in
estimating the effect of the Montreal Protocol on surface
temperature.
In one of the latest Amendments of the Montreal Protocol (the 2016 Kigali Amendment [76]), the phase-down
of hydrofluorocarbons (HFCs)—replacement chemicals of
ODSs that do not harm the ozone layer but have a large
GWP—is regulated. Without this amendment, the continued increase in atmospheric HFC concentrations would
have contributed 0.28–0.44 °C to global surface warming
by 2100. In contrast, the controls established by the Kigali
Amendment are expected to limit surface warming from
HFCs to about 0.04 °C in 2100 [77].
An unexpected slowdown in the decline of the atmospheric concentration of CFC-11 was observed after 2012
[78] and was partially caused by new emissions from eastern China (primarily the northeastern provinces of Shandong and Hebei). These emissions were likely due to new
production and use [79]. They were initially of concern
as they would delay recovery of ozone [80] and make a
small but significant contribution to global warming [81].
The emissions appear to have been eliminated [82–84] and
likely did not have a significant effect on dates of recovery
of the ozone hole [85–88]. However, if similar emissions
were to re-occur and last longer, effects on climate could
be significant.
If the production of ODSs had not been controlled by
the Montreal Protocol, biologically active UV-B radiation
causing plant damage [89] could have increased by about a
factor of five over the twenty-first century16 [90]. The ensuing harmful effects on plant growth were estimated to result
in 325–690 billion tons less carbon held in plants by the end
of this century. This reduction in carbon sequestration would
have resulted in an additional 115–235 parts per million of
CO2 in the atmosphere, causing an additional rise of global
mean surface temperature of 0.5–1.0 °C. However, these
estimates have large uncertainties and should be viewed with
caution because the “generalized plant damage action spectrum” [89] used in the calculations does not account for the
variety of plant responses across species and ecosystems.
Furthermore, experiments (summarized by Ballaré et al.
16
“World Avoided” scenarios such as the scenario discussed here
are inevitably only estimates based on the state of current knowledge.
They cannot consider possible changes in human behavior and policies that may come about when large changes in UV irradiance and
their consequences would have become more obvious in the future.
Nevertheless, these projections allow us to put the crucial benefits
that the Montreal Protocol has brought to date into perspective.
Photochemical & Photobiological Sciences (2023) 22:937–989
[91]) have not yet established whether the assumed sensitivity of plants to increases in UV-B radiation (i.e., a 3% reduction in biomass for every 10% increase in UV-B radiation for
the “reference” scenario considered by Young et al. [90]) can
be extrapolated to the very large increases in UV-B radiation
simulated in this study. For example, Young et al. [90] did
not consider that plants have protective mechanisms against
damaging amounts of UV radiation, e.g., by synthesizing
UV-absorbing compounds [e.g., 3, 92–95]. Such adaptation
would mitigate the net CO2 flux into the atmosphere. Conversely, enhanced photodegradation of organic matter under
elevated UV radiation would release additional CO2 into
the atmosphere [96]. For more details, see Box 1 of Barnes
et al. [3].
In conclusion, the studies assessed above provide further
evidence that the Montreal Protocol is not only vital for the
recovery of the ozone layer, but also for the reduction of
global warming. The Montreal Protocol is, therefore, considered to be one of the most successful international treaties
to date mitigating anthropogenic climate change.
5 Effects of recent changes in stratospheric
ozone on climate and weather
An in-depth assessment of the two-way interactions between
changes in stratospheric ozone and climate is part of the
SAP’s latest report [11]. Here, we focus on a subset of this
assessment, and emerging topics. We also highlight the
effects of Antarctic and Arctic ozone depletion on the climates of the Southern and Northern Hemisphere, respectively, and assess how these changes impact temperature and
precipitation at the Earth’s surface as well as the extent of
Antarctic sea ice and snow coverage.
5.1 Effects of Antarctic ozone depletion
on Southern Hemisphere climate
By enhancing cooling of the stratosphere, Antarctic ozone
depletion has caused a poleward shift of climate zones and
has been the primary driver of climate change in the Southern Hemisphere during summer in recent decades [97 and
Sect. 5.1.1]. An influence of stratospheric ozone changes on
sea surface temperature (SST) of the Southern Ocean may
also be expected. However, current climate models have generally not been able to reliably reproduce observed changes
in SST at high southern latitudes [98]. Recent modeling has
provided evidence that changes in atmospheric ozone during
the latter half of the twentieth century may be responsible
for about one third of the observed warming in the upper
2000 m of the Southern Ocean (30°–60° S) [99]. About
60% of this contribution can be attributed to increases in
tropospheric ozone—partly caused by increasing downward
949
transport of ozone from the stratosphere to the troposphere
and partly by enhanced production of ozone in the troposphere [100]—and the other 40% to stratospheric ozone
depletion [99].
Antarctic sea ice cover increased between 1978 and 2015
[101, 102] and has subsequently shown a general decline
with large year-to-year variability [103], which is still not
completely understood (Sect. 5.1.3). Atmosphere–ocean
interactions are intimately linked to the formation and dissipation of sea ice. However, the influence of ozone depletion on Antarctic sea ice is largely masked by other climate
processes.
5.1.1 Shifting of climate zones
The effect of stratospheric ozone depletion on the summertime large-scale atmospheric circulation in the Southern
Hemisphere has recently been confirmed and substantiated
[97]. The primary effect has been the poleward shift of the
tropospheric westerly winds over the Southern Ocean during the latter part of the twentieth century. The location of
these tropospheric winds is quantified with the Southern
Annular Mode (SAM17) index. The poleward shift of these
winds has led to a more positive state of the SAM during
summer [104–106]. This shift has affected regional temperature patterns [104] as well as precipitation in parts of
Australia and South America [107], and Antarctica [105].
Specifically, stratospheric ozone depletion led to a tendency
for more precipitation in parts of Australia, and less rain in
South America. As an example, Yook et al. [28] provide evidence that the large Antarctic ozone holes of 2020 and 2021
(Sect. 3.2)—which were likely influenced by the Australian
wildfires of early 2020 and the eruption of La Soufrière in
April 2021, respectively—contributed to anomalously strong
westerly winds over much of the Southern Ocean, anomalously cool conditions over the Antarctic plateau, anomalously warm conditions over the Antarctic peninsula, and
anomalously cool conditions over much of Australia with
flooding rains across the south-east of the continent. These
anomalies are consistent with those observed in other years
with large Antarctic ozone holes [105].
As a direct result of the Montreal Protocol, recovery of
stratospheric ozone observed since the end of the twentieth century reversed cooling trends of the Southern Hemisphere’s lower stratosphere [21, 108]. However, warming
trends observed post-2001 are about 50–75% smaller in
17
The SAM is the leading mode of Southern Hemisphere extratropical climate variability describing a seesaw of atmospheric mass
between the mid- and high-latitudes, with corresponding impacts
on the strength of the circumpolar westerly winds. A positive SAM
index corresponds to a poleward shift of the maximum wind speed,
which results in weaker-than-normal westerly winds in the southern
mid-latitudes.
13
950
magnitude than the cooling trends during the era of progressing ozone depletion. These changes in stratospheric
temperature have also halted or partially reversed the poleward shift of climate zones [97].
Projections of the future climate for the Antarctic region
under the 6th phase of the Coupled Model Intercomparison
Project (CMIP6) [109] suggest that ozone recovery over
the first half of the twenty-first century will tend to shift
the westerly jet18 equatorward during summer. This would
lead to a reversal of the changes in air and sea temperature
at the surface—as well as in precipitation and in the zonal
wind speed over Antarctica and the Southern Ocean—that
were observed during the period of progressively worsening
ozone depletion in the late twentieth century. However, this
shift in the westerly jet is countered by the effects of both
tropospheric warming and stratospheric cooling associated
with increases in GHGs. The magnitude of this effect will
depend on the GHG scenario defined by SSPs. Low-emissions scenarios (SSP1-2.6 and SSP2-4.5) tend to result in
little overall change in the jet’s position, while high-emission
scenarios (e.g., SSP5-8.5) tend to cause an overall poleward
forcing, particularly outside of the summer season. In the
second half of the twenty-first century, GHG effects dominate under all emissions scenarios, with the westerly jet
strengthened and placed further poleward than before the
ozone hole era. However, projections of how this shift will
affect weather patterns at southern mid and high latitudes
(including South America, South Africa and Australia)
are subject to the strong dependence on GHG scenario and
climate feedbacks (e.g., changes in sea ice and ocean temperatures), which may develop over the next 50 years, plus
the limited ability of models to take all these processes into
account on a regional scale.
5.1.2 Causes and consequences of the 2019/2020 “Black
Summer” fires
A topic that has emerged since our previous assessment
[9] is the role played by Antarctic ozone variability in
recent extreme weather and climatic conditions, and the
follow-on effects of these extremes for stratospheric ozone
concentrations.
From mid-2019 to early 2020, a series of devastating
wildfires occurred in Australia, particularly along parts of
the eastern coast, affecting over 10 million hectares. The
overall severity of these 2019/2020 “Black Summer” fires
was exacerbated by exceptionally hot and dry weather
18
The term “westerly jet” refers in the context to the maximum of
westerly winds (i.e., winds blowing from west to east) close to the
surface, not the jet stream in the upper troposphere. The jet’s latitude
is defined as the latitude with the largest wind speed.
13
Photochemical & Photobiological Sciences (2023) 22:937–989
conditions combined with rainfall deficits over several years.
As shown by Lim et al. [110], anomalously hot and dry conditions in subtropical eastern Australia from austral spring
to early summer are favored in years when the Antarctic
stratospheric winter vortex is weak. Weak vortex conditions
are promoted when planetary-scale (Rossby) waves disturb
and warm the Antarctic atmosphere and reduce the overall amount of stratospheric ozone depletion in spring. The
strong warming of the Antarctic stratosphere that occurred
in September 2019 is a specific case of a weak vortex that
has been linked with the 2019/2020 Black Summer fires
[25, 111–117]. Specifically, downward coupling from the
Antarctic stratosphere promoted a strong negative phase of
the tropospheric SAM at mid-latitudes in summer, which
reduced precipitation over Australia and further exacerbated
fire conditions [115].
While the fires were mainly promoted by the weak polar
vortex, the reduced ozone depletion resulting from the weak
vortex may have been an exacerbating factor. This connection was studied by Jucker and Goyal [117] who found
that surface conditions were influenced by anomalously
high concentrations of ozone in the lower stratosphere that
accompanied the stratospheric warming event and delayed
the stratosphere–troposphere coupling. This suggests that
ozone recovery could further promote a seasonal delay in
stratosphere–troposphere coupling under weak vortex conditions. On the other hand, stratospheric warming events,
such as that observed in 2019, appear to be less likely in a
future climate [118] as increasing concentrations of GHGs
will cool the stratosphere.
One consequence of the Black Summer fires was that
superheated air from these fires produced large-scale
pyrocumulonimbus clouds, which forced injection of an
unprecedented amount of smoke and tropospheric air into
the lower stratosphere [119–125]. From there, this air rose
to heights of up to 35 km where it had persistent effects
across a wide latitude band for several months [123, 126].
Ozone-poor tropospheric air in the rising plume reduced
TCO by up to 100 DU locally [119, 123, 127], with impacts
on UV radiation at the Earth’s surface. The rising air also
increased mixing ratios of water vapor in the lower stratosphere at southern mid-latitudes [123, 127] where it may
have depleted ozone through enhanced heterogeneous reactions [128], although the magnitude of this effect is unclear
[129]. The plume also contained significant quantities of
black carbon aerosol and reactive gases, which affect stratospheric chemistry [31, 130–132]. Quantifying the overall
effect of the Black Summer fires on stratospheric ozone is
still the subject of ongoing research. Additional information
on the fire’s impact on stratospheric chemistry is provided in
SAP’s latest assessment [11].
Photochemical & Photobiological Sciences (2023) 22:937–989
951
5.1.3 Effects on sea ice extent and snow coverage
depletion in the second half of the twentieth century should
have caused a reduction in sea ice extent, mainly by promoting the redistribution of ocean heat content [140, 141,
143–145]. However, only a subset of leading climate models
can adequately capture the observed link between the SAM
and autumn changes in sea ice [133]. In general, current
climate models do not provide a consistent representation
of the observed long-term trends in sea ice. As concluded
by Polvani et al. [133], this appears to be the consequence
of the relatively small fraction of variance explained by the
seasonal coupling of the SAM and sea ice, which is surpassed by the larger fractions attributable to natural variations and the models’ internal variability. The effect of ozone
depletion on changes in sea ice is, therefore, still not well
understood.
Over much of the Antarctic continent, only relatively
small seasonal changes in the short-wave albedo20 of the ice
sheet occur and are primarily caused by deposition of snow
and melting at the surface [146]. Local exceptions occur
in the regions of exposed rock, which account for approximately 0.4% of the surface area of the continent. Here, varying coverage by ice, snow, and surface water can strongly
influence albedo [147]. Changes in snowfall over Antarctica
have been attributed to changes in atmospheric circulation
resulting from the depletion of ozone [148], although patterns of relative change are heterogeneous [149].
The effects of ozone depletion on temperature and air circulation over Antarctica may also change snow and ice
cover on the Antarctic continent and the extent of sea ice.
For example, interactive climate models [97, 133], which
are state-of-the-art in representing the complex interplay
between effects of transport and dynamics [134], have demonstrated that ozone depletion has influenced near-surface
winds over the Southern Ocean during summer and could,
thus, potentially affect sea ice extent. However, as discussed
below, these linkages are still not well understood. Changes
in ice or snow coverage are important because they modify
the reflectivity of the surface, which in turn changes downwelling UV radiation (Sect. 6.2).
The sea ice zone surrounding Antarctica shows strong
seasonal variability [101, 102, 135]. There has been marked
interannual variability during the last four decades, particularly in the last years, with regionally opposing patterns
of change [Chapter 2 of 63]. Antarctic sea ice expanded
between 1979 (the start of satellite measurements) and 2015,
although only in the transitional seasons. Trends in both
summer and winter were not significant. After this period
of increase, the extent of Antarctic sea ice declined dramatically during the austral springs of 2016 and 2017 [101],
reaching a record low on 1 March 2017, which was 27%
below the mean of annual minima calculated for 1978–2016.
However, a partial recovery was observed between 2017 and
2021 (https://climate.nasa.gov/ask-nasa-climate/2861/arctic-and-antarctic-sea-ice-how-are-they-different/). Several
studies examined the reasons for this recovery [136–138];
however, none of these studies found robust evidence that
trends or variations in stratospheric ozone contributed to
this phenomenon.
As discussed in Sect. 5.1, the effect of stratospheric ozone
depletion on temperatures at the surface of the Southern
Hemisphere is primarily mediated by changes in the SAM
during summer. Observational studies have shown that the
seasonal response to trends in the SAM has resulted in the
cooling of the SST around Antarctica in autumn, which
should have promoted an overall increase in sea ice extent
in that season, consistent with observations between 1979
and 2015 [139–141]. Furthermore, ozone depletion has been
linked to a reduction in downwelling long-wave19 radiation. This reduction would also cool the Southern Ocean
[142, 143]. In contrast to these studies, results from stateof-the-art earth-system models clearly indicate that ozone
19
Long-wave radiation is electromagnetic radiation with wavelengths between 3 and 100 μm that is emitted from the Earth and its
atmosphere in the form of thermal radiation. Long-wave radiation
contrasts with short-wave radiation with wavelengths between ~ 0.3
and ~ 3 μm originating from the Sun.
5.2 Associations between Arctic stratospheric
ozone losses and the climate of the Northern
Hemisphere
Years with a strong Arctic polar vortex and associated significant stratospheric ozone depletion have been linked to
widespread climate anomalies across the Northern Hemisphere based on targeted model experiments with CCMs
[150]. As an example, the exceptionally large ozone depletion that occurred in March–April 2020 (Sect. 3.3) not
only led to record-breaking increases in Arctic solar UV
radiation (Sect. 7.1.2) but also affected weather patterns
in the Northern Hemisphere during spring. Specifically, it
helped to keep the Arctic Oscillation (or AO21) in a recordhigh positive state through April [36], thus contributing to
abnormally high temperatures across Asia and Europe [151].
20
Albedo is the proportion of the incident radiation that is reflected
by a surface. Short-wave albedo refers to the fraction of the total incident solar irradiance in the wavelength range of ~ 0.3–3 μm that is
reflected by the Earth’s surface. Albedo may also refer to the reflectivity in a certain wavelength range, such as the UV range.
21
The Arctic Oscillation (AO) or Northern Annular Mode (NAM)
is analogous to the Southern Annular Mode (SAM) and characterizes
the pattern of winds circulating around the Arctic. When the AO is in
its positive phase, a ring of strong winds circulating the North Pole
acts to confine colder air in the polar regions.
13
952
Furthermore, loss of stratospheric ozone modified circulation patterns of winds around the Arctic, thereby affecting
the stability of the upper troposphere in the Siberian sector
of the Arctic. In turn, this led to more high-level clouds that
enhanced downwelling long-wave (thermal) radiation [152].
The associated anomalous warming of the surface in April
2020 was further amplified by a reduction in albedo caused
by melting of snow and sea ice. Monthly anomalies (relative to the 1981–2010 climatology) in air temperature of up
to + 6 °C were observed over Siberia from January through
May 2020 [153]. The temperature in the Siberian town of
Verhojansk (68° N, 133° E) set a new record of 38 °C on
20 June 2020, which is the highest temperature ever documented near the Arctic Circle. Depletion of stratospheric
ozone over the Arctic in March may cause reductions in
the sea ice concentration and the sea ice thickness over the
Arctic Ocean north of Siberia from spring to summer [154].
The unprecedented depletion of Arctic ozone in the spring
of 2020 contrasts with the boreal winter of 2020/2021, when
a major sudden stratospheric warming (SSW) occurred on
5 January 2021 [42, 43] and limited overall ozone loss
(Sect. 3.3). During an SSW event, the westerly winds of the
wintertime polar stratosphere decelerate and temperatures in
the polar stratosphere rapidly increase [155]. The 2021 SSW
event warmed the lower stratosphere, interrupted the catalytic cycles associated with ozone depletion [47], and also
affected the polar atmospheric circulation from the upper
stratosphere to the surface for six weeks after the event. During this period, surface temperatures were anomalously high
over Greenland and the Canadian Arctic and anomalously
low over Europe, northern Asia, and the United States, with
a cold air outbreak first occurring over Eurasia in January
and then over North America in the first two weeks of February [156]. SSWs generally increase the likelihood of such
weather anomalies [157]; however, it is still unclear to what
degree the cold weather events in early 2021 were linked to
the SSW on 5 January 2021. There is some evidence that the
cold outbreak in Siberia on 22–24 January 2021 was associated with the SSW [158]. However, simulations with a climate model did not find evidence that this SSW event caused
or influenced the record-breaking cold in North America
during February 2021.
Precipitation in Central China in April–May has been
linked to Arctic stratospheric ozone changes in February–March by combining observations, reanalysis data, and
a CCM [159]. Specifically, positive Arctic ozone anomalies
enhance precipitation in central China and negative anomalies reduce precipitation. Another study, using the same
CCM, demonstrated a negative relationship between Arctic
ozone anomalies in March and surface temperature anomalies in central Russia and, a weaker positive relationship in
southern Asia [160]. Furthermore, variations in precipitation
occurring during April in the northwestern United States
13
Photochemical & Photobiological Sciences (2023) 22:937–989
(mainly the states of Washington and Oregon) are strongly
linked to changes in Arctic stratospheric ozone during
March [161]. Specifically, higher-than-normal Arctic ozone
concentrations in March lead to less precipitation in April
and vice versa.
Despite these advances in the understanding, assessing
linkages between Arctic ozone depletion and weather in the
Northern Hemisphere remains difficult and is subject to large
uncertainties. It is anticipated that future studies will refine
the conclusions summarized above.
6 Factors other than ozone affecting UV
radiation
Solar UV-B radiation at the Earth’s surface is mostly controlled by the height of the Sun above the horizon (i.e., the
solar elevation22); TCO; clouds; aerosols; the reflectivity of
the surface, also called albedo; and altitude. Less important factors include: the vertical distribution of ozone in the
atmosphere (i.e., the ozone profile) for fixed TCO; other
trace gases such as sulfur dioxide (SO2) and nitrogen dioxide
(NO2); seasonal changes in the Earth–Sun distance; changes
in solar activity, which influence both stratospheric ozone
concentrations and the UV-B irradiance at the top of the
atmosphere; topography; and volcanic eruptions. Except for
determinants related to the Sun and volcanic activity, all
these factors are influenced by human activities—such as the
release of GHGs and air pollutants—and are coupled with
changes in the climate. For example, higher temperatures
will lead to less sea ice in the Arctic, which will in turn
reduce surface reflectivity and UV radiation at or above the
surface. The effects of these factors have been described at
length in previous assessments [9, 162, 163]. No studies
published in the last four years provide new insights into
the effect of clouds on UV radiation. We, therefore, focus in
the following sections on new understandings into the roles
of aerosols, albedo, solar activity, volcanic eruptions, and
climate interactions on UV radiation.
6.1 Aerosols
Natural and anthropogenic aerosols (solid and liquid particles suspended in the atmosphere) play a major role in
controlling the intensity of UV radiation at the Earth’s
surface. Although effects of aerosols have been discussed
in numerous studies, the magnitude of these effects is still
22
The position of the Sun in the sky is typically either described by
the solar elevation, which is counted from the horizon, or the solar
zenith angle (SZA), which is counted from the zenith (the imaginary
point directly above a particular location). The solar elevation can be
calculated as 90° – SZA.
Photochemical & Photobiological Sciences (2023) 22:937–989
uncertain. The attenuation of surface UV radiation by aerosols depends on their amount, as measured by aerosol optical
depth (AOD), and on their efficiency of absorption, as discussed at length in our last assessments [9, 162]. To quantify
these effects further, Campanelli et al. [164] analyzed optical properties of aerosols and spectral irradiance in Rome,
Italy, and correlated the variability of the UVI (adjusted for
variations in TCO) with the AOD at 340 nm for two groups
of either strongly or weakly absorbing aerosols. Absorption
for the two classes was quantified with the single scattering
albedo23 (SSA). For strongly absorbing aerosols (SSA < 0.9),
an increase of the AOD by one unit resulted in a decrease
of the UVI by 2.7 units (about 30%) for a solar zenith angle
(SZA) of 30° and by 1.65 units (about 25%) for a SZA
of 40°. For less absorbing aerosols (SSA > 0.9), the UVI
decreased only by one unit (about 12%) per unit of AOD
increase for both SZAs. The study illustrates the importance
of the absorption properties of aerosols.
The paucity of measurements of the properties of aerosols (including the SSA) in the UV-B range [9] hampers
our ability to accurately assess the effects of aerosols on a
global scale as well as for urban regions with a diverse mix
of aerosol types [5]. Global networks, such as the Aerosol Robotic Network (AERONET), which measure AOD
and other aerosol properties, do not perform observations
at UV-B wavelengths. While the technology for measuring
AOD and SSA in the UV-B range exists and has been tested
at a few sites [165–169], there are at present no reliable data
to assess aerosol properties in this critical wavelength range
on a global scale. However, the European Brewer Network
(EUBREWNET) [170] has recently started to provide AOD
in the wavelength range from 306 to 320 nm [167] and a
preliminary analysis confirms the good quality of the data.
It is anticipated that this network will expand globally.
In areas with elevated levels of air pollution and small
variability in TCO, the attenuation of solar UV radiation
under cloudless skies is mainly controlled by aerosols. In
such areas, abatement of air pollution can lead to increases
in the intensity of UV radiation towards levels that would
normally occur in unpolluted areas at similar latitudes
and altitudes. An example of this is the observed increase
of ~ 25% in UVI over Mexico City between 2000 and 2019,
which was attributed to reductions in pollutants; in order
of importance, aerosols, tropospheric ozone, NO2, and SO2
[171]. Because of high historical levels of air pollution
in Mexico City, the UVI under cloud-free conditions was
953
Fig. 5 Monthly average noontime UVI in the Mexico City Metropolitan Area (black line) ± 1 standard deviation (blue shading), and linear
fit (red line) to average data. Reprinted from Ipiña et al. [171] with
permission from the American Chemical Society, Copyright © 2021
lower by ~ 40% in 2000 and ~ 25% in 2019 relative to values
expected for an unpolluted clear atmosphere. Monthly averages of the daily maximum UVI from the 11 stations distributed across the Mexico City Metropolitan Area considered
in this study show a clear upward trend of 0.9% per year
between 2000 and 2019, and an overall increase in monthly
maximum UVI of 1.5 over the two decades (Fig. 5). Since
2016, the rate of increase is greater, possibly reflecting more
aggressive measures in reducing air pollutants. Human
health benefits resulting from the decrease in air pollution
[5, 172] outweigh risks—such as the potential increase in
skin cancer incidence—stemming from the gradual return
of UV radiation intensities to more natural levels prevailing
at unpolluted areas24.
The effects of air quality measures implemented in
Mexico City may help to project changes in UV radiation
for regions that are currently still affected by heavy smog,
such as South and East Asia [9, 173]. Finally, the study for
Mexico City also confirmed earlier findings [e.g., 174] that
the UVI at the surface of heavily polluted areas cannot be
reliably estimated from satellite observations, emphasizing
the importance of ground-based measurements. A similar
finding was reported by Roshan et al. [175] for the city of
Doha, Qatar, when extreme dust storms resulted in a measured UVI of 6–7 compared to a UVI of 10–11 estimated by
the OMI satellite on the same days.
23
The AOD is the sum of the aerosol scattering optical depth τsca
and the absorption optical depth τabs, which quantifies the attenuation
of the direct solar beam due to scattering and absorption of photons:
τ = τsca + τabs. Instead of specifying τ and τabs, the single scattering
albedo (SSA) is often reported instead: SSA = τsca / (τsca + τabs) = τsca /
τ, resulting in τabs = (1 − SSA) × τ. A decrease in SSA, therefore, corresponds to an increase in absorption of radiation.
24
The global number of deaths from air pollution (particulate matter and gases such as tropospheric ozone and nitrous oxides) has been
estimated at 4.2 million per year [5]. In comparison, the number of
deaths from skin cancer was about 120,000 in 2020 (https://gco.iarc.
fr/today/fact-sheets-cancers, accessed 13 November 2022).
13
954
The effect of increased aerosols and tropospheric ozone
on surface UV radiation during the biomass burning25 season
in Pretoria, South Africa, was investigated by du Preez et al.
[176]. The simulations included different scenarios with and
without increased levels of aerosols and tropospheric ozone
from biomass burning. For cloudless days during the height
of the biomass burning period in September, aerosols and
tropospheric ozone reduced the noontime UVI by 13% and
1%, respectively, demonstrating that changes in the UVI
were dominated by the effects from aerosols.
Smog from the Black Summer wildfires in Australia
(Sect. 5.1.2) led to extreme air pollution and low visibility. However, even during days with a visibility of less than
5 km, the intensity of UV radiation may have still been
harmful to human health. For example, on 10 December
2019, the visibility near Sydney, Australia, dropped to about
1 km around noon. Despite this low visibility, the cumulative erythemal UV dose measured at this location over
a one-hour period at noon was still more than 4 SED26 or
about 46% of the one-hour dose measured on the cloud and
haze-free day of 27 November 2019. During the eight hours
between early morning and late afternoon the dose on 10
December 2019 was 17 SED. The corresponding dose of 27
November was 48 SED [180]. These UV doses far exceed
the maximum daily UV dose recommended by ICNIRP for
outdoor workers [181]. While most people stayed indoors
during the fires because the air pollution was so extreme,
emergency workers, who had to be outside despite adverse
conditions, may have been exposed to UV radiation levels
harmful to human health, potentially without being aware of
it and without applying appropriate sun protection measures.
Despite increases of aerosols in specific regions (e.g.,
from bushfires, burning of biomass or dust storms), over
most populated areas of the globe, there is a general decrease
in aerosols. Trends of aerosol optical and chemical properties on global and regional scales have been reported from
observations with several ground-based networks [182].
Most of the properties related to loading of aerosols exhibit
negative trends in the period 2000–2014 in regions covered
by observations, both at the surface and in the total atmospheric column. Significant decreases in AOD were found
in areas with intense anthropogenic activity (Europe, North
America, South America, North Africa and Asia), ranging
25
Biomass burning is the burning of living and dead vegetation. It
generally includes the human-initiated burning of vegetation for land
clearing and land-use change as well as natural, lightning-induced
fires.
26
The standard erythemal dose (SED) is a measure of cumulative
erythemal UV radiation [177]. One SED is equivalent to an erythemally effective radiant exposure of 100 J m−2. Two SED may lead to
erythema in individuals with freckled pale skin (Skin Type I, defined
by the Fitzpatrick scale [178]). Longer exposure times are required
for individuals with darker skin [179].
13
Photochemical & Photobiological Sciences (2023) 22:937–989
from − 1.2% per year to − 3.1% per year. These data were
used to validate various aerosol models (six AeroCom27
phase III models, four CMIP6 models and the CAMS28
reanalysis dataset) showing good agreement in the AOD
trends. When these models were used to estimate the global
AOD trend by filling the gaps in regions not covered by
observations, a global increase in AOD of about 0.2% per
year between 2000 and 2014 was found, primarily caused
by an increase in the loads of organic aerosols, sulfate, and
black carbon. These findings highlight differences between
regional and global effects of aerosols on UV radiation,
which must be considered, especially when projecting into
the future.
In a modeling study [183] exploring China’s future
anthropogenic emission pathways, it was projected that
emissions of major air pollutants (i.e., SO2, NOx, PM2.5
aerosols, and non-methane volatile organic compounds) in
China will be lower by 34–66% in 2030 and by 58–87%
in 2050 compared to 2015. These estimates were derived
by considering a combination of strong low-carbon and air
pollution control policies. A second study [184] investigated
the evolution of different types of aerosols over the EuroMediterranean region between 1971–2000 and 2021–2050
according to three different scenarios representing a wide
range of possible future pathways. The study showed a
decrease in AOD of between 30 and 40% over Europe,
mainly from decreasing emissions of sulfur dioxide. However, these reductions are partly (~ 30%) compensated by
increases in the optical depth from nitrate and ammonium
particles.
Attenuation of UV radiation by aerosols can sometimes
also mask the effect of “ozone mini-holes” (defined as a
synoptic-scale29 region with strongly decreased TCO resulting from dynamical processes [185]) that would otherwise
lead to increases in UV radiation. One example is an event
that occurred in Athens, Greece, during 8 days in May 2020
[186]. On 15 May 2020, TCO was 43 DU (or more than 2
standard deviations) below the climatological mean, which
would have normally led to an increase in the UVI by ~ 29%.
However, the AOD on this day was 0.31 (47%) higher than
the climatological mean due to the intrusion of Saharan dust,
and measured UVIs agreed to within ~ 2% with the climatological mean. Hence the opposing effects of low TCO and
high AOD nearly canceled each other. This study highlights
the important role of aerosols in modifying the effects of
changes in TCO on surface UV-B radiation. There is some
27
Aerosol Comparisons between Observations and Models (https://
aerocom.met.no/).
28
Copernicus Atmosphere Monitoring Service (https://atmosphere.
copernicus.eu/).
29
In meteorology, synoptic scale refers to a high- or low-pressure
area with a horizontal length scale of the order of 1000 km or more.
955
75
25
70
20
65
15
60
10
55
5
50
2005
2010
2015
Distance to ice edge [km]
Spectral irradiance between
–2
337.5 and 342.5 nm [ W cm ]
Photochemical & Photobiological Sciences (2023) 22:937–989
0
Year
Fig. 6 Comparison of monthly mean spectral irradiance between
337.5 and 342.5 nm for March (left axis) at Arrival Heights, Antarctica, and approximate distance of the outer edge of the land-fast
ice from Arrival Heights (right axis) during March. Distance data
are based on Fig. 3 of Kim et al. [189]. The vertical extension of the
blue bars indicates the variability of the distance within the month of
March. A distance of zero km from the ice edge means that McMurdo
Sound, 1 km west of Arrival Heights, was free of ice. In years when
the ice edge was far from the station and the ocean surrounding the
station was covered by sea ice, the albedo was greatly enhanced
and UV radiation in these years tended to be higher compared to
years when McMurdo Sound adjacent to the station was free of ice.
Adapted from Bernhard and Stierle [190]
evidence that the weather pattern that led to the transport of
dust from Africa towards Athens was also responsible for
the occurrence of the ozone mini-hole and the low TCO over
Athens that ensued.
6.2 Surface reflectivity
Changes in the reflectivity of the Earth’s surface (both
land and ocean) can change the downwelling UV radiation
because radiation that is reflected upward by the surface
may subsequently be scattered downward by air molecules,
aerosols, and cloud droplets. Topography can modify the
reflectivity resulting in complex effects on UV radiation, as
for example in narrow valleys with snow covered slopes. The
largest effect of surface reflectivity occurs in areas with variable snow and ice cover because of the large difference in the
albedo of bare and snow/ice-covered ground. This variability is often linked to climate change. For example, because
of the warming of the Arctic, the start date of the spring
snow melt at Ny-Ålesund (79° N), Svalbard, has advanced
by three days per decade over the last 40 years [187], so
now begins about two weeks earlier than in the early 1980s.
Figure 6 illustrates the effect of surface albedo
on UV radiation by comparing UV irradiance in the
337.5–342.5 nm range measured at Arrival Heights (78° S),
Antarctica, with the extent of land-fast ice—defined as sea
ice fixed in place by attachment to land, glaciers, grounded
icebergs, or ice shelves—covering McMurdo Sound 1 km
west of Arrival Heights. In 2000, a mega-iceberg calved
from the Ross Ice Shelf, became temporarily trapped, and
persisted in the entrance to McMurdo Sound for five years
[188]. The tabular iceberg interrupted the normal movement
of sea ice, resulting in McMurdo Sound remaining covered
by ice with high albedo until April in some years [189].
As a consequence, UV irradiance was elevated in March
between 2001 and 2007 when the ice edge was more than
13 km away from McMurdo, while less UV radiation was
observed between 2011 and 2015 when McMurdo Sound
was free from land-fast ice [190]. Since similar data are not
available before 2000 and after 2016, sea ice cannot be correlated with UV radiation over a longer time period at this
location. However, Kim et al. [189] reported that the dates
of the retreat of land-fast ice in McMurdo Sound have not
changed over the last 37 years except for years affected by
mega-icebergs.
6.3 Solar activity
Variability in the solar activity can indirectly affect UV
radiation at the Earth’s surface through changes induced in
atmospheric ozone, particularly in the stratosphere. These
changes in ozone are caused by two different mechanisms,
which are both related to the 11-year variability of solar
activity. One mechanism is mediated through photochemical processes in the upper atmosphere that are modified by
changes in solar UV-C (100–280 nm) radiation. The other
process is driven by changes in the rate of energetic particle30 precipitation (EPP), which mainly affect ozone over the
polar regions [191, 192].
The increase in emissions of solar UV-C radiation
between the minimum and maximum of the solar cycle leads
to increases in ozone concentrations in the upper stratosphere
(altitude of 30–60 km) and decreases in the lower stratosphere (15–30 km), mainly at lower latitudes [193]. Using a
CCM, Xiao et al. [194] estimated that for a 5% (10%) increase
in solar output in the spectral range of 200–370 nm, the globally averaged ozone increases by up to 4.5% (9.0%) in the
upper stratosphere, and decreases by up to 1.5% (3.3%) in the
lower stratosphere. It was further noticed that the response
of ozone to the variability of UV-C radiation during a solar
cycle is non-linear, confirming earlier results [195].
Our previous assessment [9] discussed the effects of
reduced solar activity in the future (e.g., from a Grand Solar
Minimum31) on UV-B radiation received at the Earth’s surface.
30
Energetic particles considered here are highly energetic electrons,
protons, neutrons, and ions that are accelerated into the atmosphere
through various heliophysical and geomagnetic processes. They enter
the atmosphere mainly in the geomagnetic polar regions (https://lasp.
colorado.edu/home/mag/research/energetic-particle-precipitation).
31
Grand solar minima are defined as periods when several solar
cycles exhibit lesser than average activity for decades or centuries.
13
956
Based on the work by Arsenovic et al. [196], we concluded that
UV-B radiation at the top of the atmosphere would decrease
slightly due to weaker emission from the Sun; however, the
reduced solar activity would also lead to decreases of ozone
production in the stratosphere, resulting in an overall increase
of UV-B radiation at the surface. This conclusion is still valid.
Solar activity has recently shown a declining tendency, suggesting that the Sun has entered into a modern Grand Solar
Minimum period, from about 2020 to 2053, which would lead
to a significant reduction of the solar magnetic field and magnetic activity by about 70%, similar to the Maunder minimum
that occurred in the period 1645–1710 [197]. The influence
of such reductions in total solar irradiance (TSI32) on surface
temperatures was investigated using a climate model run under
the RCP 8.5 scenario, which predicted a decrease in the global
average temperature for the second half of the twenty-first century of 0.13 °C due to atmospheric effects of the upcoming
Grand Solar Minimum [198]. Simulations by Arsenovic et al.
[196], which were based on the RCP 4.5 GHG scenario, estimated that a stronger solar minimum with reduction in TSI of
0.48% would only compensate for about 15% of GHG-induced
warming by 2100. Hence, the estimated decreases in temperature by 2100 due to reduced solar activity are small compared
to the projected increases due to GHG emissions. Therefore,
the reduction of solar irradiance during a possible Grand Solar
Minimum would only partly offset the anthropogenic change
in climate caused by continuing GHG emissions.
The upcoming maximum of Solar Cycle 25 is expected to
be weaker than the current Cycle 24, which was the weakest
in at least the past 100 years [199, 200]; however, the uncertainty of this prediction is large. Model results [199] estimated
a deep extended solar activity minimum for 2019–2021, and
a weak solar activity maximum in 2024–2025. This modeling study is based on analysis of magnetograms that contain
information on the evolution of magnetic fields on the solar
surface, allowing forecasting of the solar activity in the future.
The reduced activity in the period of the solar maximum will
lead to less photochemical production of stratospheric ozone
at low latitudes, but also to reduced polar ozone destruction
due to fewer energetic particles.
Although none of the studies discussed above addressed
effects on surface UV-B radiation, the upcoming weaker
solar activity period would lead to decreases in stratospheric
ozone and consequently to increases in UV-B radiation at
the surface, despite the reduced solar irradiance entering the
Earth’s atmosphere. This effect has not been considered in
the projections of UV radiation described in Sect. 8.
32
Total Solar Irradiance (TSI) is the solar radiative power per area
integrated over all wavelengths that is incident on the Earth’s upper
atmosphere.
13
Photochemical & Photobiological Sciences (2023) 22:937–989
6.4 Volcanic eruptions
Throughout the Earth’s history, major volcanic eruptions or
impacts of meteors have perturbed the climate, affected the
stratosphere, and caused regional and global environmental
disasters. Global effects are mainly caused by the reflection
of incoming solar radiation by the aerosol layer that forms in
the stratosphere after a large volcanic eruption, but also by
the destruction of stratospheric ozone involving heterogeneous chemical reactions on the surfaces of volcanic aerosols
in the presence of halogens [201–203]. Volcanic aerosols
are dispersed zonally to other latitudes and can persist for
several years, resulting in cooling of the troposphere. Such
eruptions can either reduce solar UV-B irradiance at the
Earth’s surface through scattering of radiation back to space
or increase it through reduced absorption by the depleted
ozone layer. The magnitude of these effects depends on the
strength of the eruption and on the amounts of aerosols and
halogenated compounds involved. Large tropical volcanos in
the last ~ 200 years, e.g., Mt. Pinatubo in 1991 and Mt. Tambora in 1815, have caused globally averaged cooling of 0.3
and 0.7 °C at the Earth’s surface, respectively [204]. Conversely, stratospheric aerosols from Mt. Pinatubo warmed
the lower tropical stratosphere by up to 4 °C in the 2–3 years
following the eruption [205].
Recent studies used chemistry–climate models to investigate the effects of different amounts of SO2 and halogens
injected into the tropical stratosphere by volcanic eruptions
[204, 206, 207]. Brenna et al. [206] assumed an explosive
eruption at 14° N, rich in sulfur, chlorine, and bromine compounds, occurring during pre-industrial times. The assumed
amount of SO2 injected represents the average of 28 historical volcanic eruptions in the Central American Volcanic
Arc (CAVA; extending parallel to the Pacific coastline from
Mexico to Panama), comparable in magnitude with the Mt.
Pinatubo eruption. However, the amount of bromine and
chlorine deposited in the stratosphere was assumed to be
much larger than the amount estimated for Mt. Pinatubo.
(The Mt. Pinatubo eruption was unusual because it occurred
at a time when the Philippines were also inundated by a
typhoon. Water droplets from the storm likely adsorbed
halogen compounds in the plume and prevented them from
reaching the stratosphere [208, 209].) The ozone depletion
calculated by Brenna et al. [206] led to increases in the clearsky summertime UVI of more than 50% in the NH during
the first two years after the eruption. Maximum increases in
the UVI were modeled to exceed 7 units in the NH tropics
and subtropics, and peak at 4 units in the NH mid-latitudes.
Much of the mid-latitudes would have experienced a UVI
above 15, which is similar to present-day peak values in the
tropics [210]. This simulation was based on the injection
of large amounts of halogens, which are thought to be representative for volcanic eruptions in the CAVA. Simulated
Photochemical & Photobiological Sciences (2023) 22:937–989
increases in UV radiation are therefore much larger than
those observed after the eruption of Mt. Pinatubo.
Another modeling study [204] investigated the effect on
the atmosphere of the eruption of a super volcano like Toba,
which erupted 74,000 years ago. It has been estimated that
Toba injected 100 times more SO2 into the stratosphere
than Mt. Pinatubo. According to this study33, such an event
could lead to the collapse of the ozone layer in the tropics with ~ 50% reduction in TCO, which would increase the
daily maximum UVI by more than a factor of two. Even
with one fifth of the injected SO2 amount, ozone depletion
in the tropics would be similar to that currently occurring in
Antarctica and would last for nearly a year.
These studies show that massive but rare volcanic eruptions can lead to severe depletion of stratospheric ozone,
changes in atmospheric circulation and temperature patterns, and large increases in UV-B radiation at the Earth’s
surface. These increases can by far exceed those associated
with ozone depletion from ODSs in the 1980s and 1990s as
well as the expected rise in UV-B radiation to more natural
levels over urban regions that may occur when measures to
reduce air pollution are implemented (Sect. 6.1).
6.5 Climate change
In the absence of changes in the TCO, climate-changeinduced trends in the properties of clouds, atmospheric
aerosols and surface albedo have the potential to strongly
influence the long-term behavior of UV radiation at the
Earth’s surface.
The optical properties of clouds, aerosols, and surface
albedo, and the interactions between these components, are
active areas of research because of their importance in the
radiative balance at the surface. Global warming is expected
to influence cloudiness because of the atmosphere’s ability
to hold more water as temperatures increase [211]. However, patterns of change in cloud cover, height, and optical
depth are difficult to assess because of the inherent internal
variability in regional climate forcing combined with the
short length of available climate data records. The physical
understanding of cloud processes continues to advance. For
example, the better understanding of the microphysics of
33
The study did not include the chemical impact of halogen compounds, as there is no reliable information on their emissions from
Toba. The modeled effect on ozone mainly occurs because absorption
by SO2 and scattering by the aerosol layer reduces the flux of solar
UV-C radiation reaching the lower stratosphere. Solar UV-C radiation with wavelengths shorter than 242 nm initiates the formation of
ozone in the stratosphere because it leads to the photolysis of oxygen
molecules (O2). The resulting oxygen atoms react with O2 to form
ozone (O3). Less UV-C flux below the aerosol layer therefore leads to
less ozone production. If ozone loss by halogen compounds had been
included also, the modeled ozone decline and increase in UV-B radiation at the Earth’s surface would have been even larger.
957
supercooled liquid water has reduced the bias in the modeled
short-wave cloud radiative effect over the Southern Ocean
[212]. Climate models also continue to improve in their representation of aerosols, which cool the lower troposphere
and counter some of the warming resulting from GHGs
[63]. Reductions in air pollution have generally occurred in
Europe and North America as the result of regulations; however, economic growth has caused large regional increases
in aerosol emissions in Asia and Africa [213]. Interactions
between aerosols and clouds remain the largest uncertainty
in climate projections. Changing patterns of coverage of the
surface with snow, ice, and vegetation under global warming
are also relevant to surface UV irradiance, with observed
darkening of the Arctic surface over 2000–2019 attributed
to summertime loss of sea ice, while mixed trends in albedo
have occurred over this period in Antarctica [214] (see also
Sect. 6.2).
The complexity in accurately accounting for all relevant
processes, particularly on small scales where observations
are influenced by local effects (e.g., UV enhancement under
broken clouds), limits the ability to attribute trends in UV
irradiance to specific climate change effects. However, several recent studies have quantified local-scale influences,
with examples provided below.
The occurrence of cloud-free conditions is very important for total UV exposure. Atmospheric blocking systems,
which are large-scale patterns of stationary atmospheric
pressure fields that “block” or redirect migratory cyclones or
anti-cyclones, can lead to prolonged periods of clear skies at
mid and high latitudes. In a blocking event, a high-pressure
weather system can persist for days or even weeks over some
geographical regions, inhibiting cloud formation and causing
moisture in the westerly zonal flow to be deflected around it.
Hence, clouds are often more persistent than usual outside
regions with high pressure resulting in lower UV irradiance
at the Earth’s surface. A recent example where surface UV
radiation was exceptionally affected by atmospheric blocking occurred during May–July 2018 in Norway and Finland
[215]. The monthly mean noontime UVI was 20–40% above
the long-term mean as a direct result of decreased cloud cover.
For example at Sodankylä (67° N), the mean temperature in
July 2018 was 5.6 °C above the 1981–2010 average for the
same month and the duration of sunshine in 2018 was 405 h,
exceeding the 1981–2010 average of 245 h by 65%. This particular event was associated with a record heat-wave in central
and northern Europe [216]. Recent studies examining trends
and variability in atmospheric blocking at high latitudes have
found mixed patterns of change, with regional shifts in trends
in the Antarctic Peninsula region over the satellite era [217],
and no significant trends over Greenland [218]. For highemissions SSP scenarios, a clear decrease in future blocking
over Greenland and the north Pacific was found, but seasonal
and regional projections are generally unclear [219].
13
(a)
4
3
2
Daily Maximum UV Index
South Pole, Antarctica (90 S)
1991 – 2017
2018
2019
2020
2021
1
0
1-Sep
Daily Maximum UV Index
It has been known for decades that changes in tropopause
height are inversely linked to changes in TCO [220, 221]. If
the tropopause is shifted up, some lower stratospheric ozone
is horizontally transported to surrounding regions with
lower tropopause height. The result is a decrease of TCO in
areas where the tropopause is elevated [221]. Furthermore,
mid-latitude regions with elevated tropopause may also be
influenced by the advection of stratospheric ozone-poor air
masses from lower latitudes (ozone mini-holes) [185, 222].
In a new study, Fountoulakis et al. [223] quantified the
effect of changes in the geopotential height (GPH) at 250 hPa
(a quantity similar to tropopause height) on TCO and spectral
irradiances at 307.5 and 324 nm at three locations across Italy:
Aosta (46° N), Rome (42° N), and Lampedusa (36° N). Statistically significant anti-correlations were found between GPH and
monthly anomalies in TCO for all locations and months. Conversely, positive correlations between GPH and monthly anomalies in spectral irradiance at 307.5 nm were detected for most
months. The study makes a strong case that increases in GPH
or tropopause height that are expected from the warming of
the troposphere due to climate change [224, 225] would reduce
TCO and subsequently lead to increases in UV-B radiation.
Additional effects of climate change on TCO and the
vertical distribution of ozone in the atmosphere—such as
the expected strengthening of the Brewer-Dobson circulation, unexpected declines in lower stratospheric ozone in
the extratropics [226], and the dependence of TCO on GHG
scenarios (Sect. 3.5)—are discussed in great detail in SAP’s
latest report [11] and are therefore not addressed here.
Photochemical & Photobiological Sciences (2023) 22:937–989
Daily Maximum UV Index
958
8
1-Oct
1-Nov
1-Dec
1-Jan
(b) Arrival Heights, Antarctica (78 S, 167 E)
7
1989 – 2017
2018
2019
2020
2021
6
5
4
3
2
1
0
1-Sep
16
14
1-Oct
1-Nov
1-Dec
1-Jan
(c) Palmer Station, Antarctica (65 S, 64 W)
12
10
8
6
1990 – 2017
2018
2019
2020
2021
4
2
0
1-Sep
1-Oct
1-Nov
1-Dec
1-Jan
Month
7 Variability in UV radiation and trends
from observations
This section assesses observed variations in UV radiation on
various time scales as well as long-term trends in the UVI
observed by ground-based and space-borne instruments over
several decades.
7.1 Variations in UV radiation with time
and altitude
Year-to-year and seasonal variability in UV radiation is
mainly controlled by variations in the TCO, cloud cover,
and aerosols. For example, TCO at mid-latitudes is higher
in the spring and lower in the autumn. As a result, the UVI
near the autumn equinox can exceed that at the spring equinox by nearly a factor of two for matching SZAs [227]. The
effect from ozone is most pronounced at high latitudes of
the Southern Hemisphere during spring but variability in
stratospheric ozone in the Arctic has also led to larger variability in UV radiation at northern high latitudes in recent
13
Fig. 7 Daily maximum UVI measured at a the South Pole, b Arrival
Heights, and c Palmer Station in 2018 (green), 2019 (yellow), 2020
(red), and 2021 (blue) compared with the average (white line) and the
range (gray shading) of daily maximum observations of the years indicated in the legends. The UVI was calculated from spectra measured by
SUV-100 spectroradiometers. Up to 2009, the instruments were part of
the NSF UV Monitoring Network [228] and they are now a node in the
NOAA Antarctic UV Monitoring Network (https://gml.noaa.gov/grad/
antuv/). Consistent data processing methods were applied for all years
[190, 229]
years during the late winter and early spring season. Both
regions are discussed in the following sections.
7.1.1 Temporal variations of UV radiation in Antarctica
We reported in our previous assessment [9] that the variability of UV-B radiation in Antarctica observed between
2014 and 2017 was very large, with near record-high UVIs
observed at the South Pole in spring 2015, and well below
average values in spring 2016. Variability during the period
discussed in this report (2018–2021) was equally large,
Photochemical & Photobiological Sciences (2023) 22:937–989
despite evidence that stratospheric ozone concentrations
over Antarctica are now recovering (Sect. 3.2).
Figure 7 shows the daily maximum UVI observed at three
Antarctic stations (South Pole (90° S), Arrival Heights (78° S),
and Palmer (65° S)) for September–December, the months
most affected by the ozone hole. Observations in 2018, 2019,
2020, and 2021 were compared with the average and range
of measurements between ~ 1990 and 2017. UVIs in October
2018 were well above the long-term mean and approached
historical maxima at the South Pole but remained within the
range of typical variability at the other two sites. Conversely,
unusually low UVIs were observed at the South Pole and
Arrival Heights in spring 2019 due to a record-high TCO during this period. Between October and mid-November 2019,
the UVI at the South Pole was at the minimum of the historical (1991–2017) range and remained close to this minimum
between mid-November 2019 and January 2020. At Arrival
Heights, the UVI in 2019 was close to the minimum between
September and mid-November, and stayed below the longterm mean until mid-December, except for two short periods.
In contrast to 2019, near record-high UVI maxima were
observed in spring 2020 and 2021 because of large and persistent Antarctic ozone holes in these years (Sect. 3.2). In both
years, the UVI at the South Pole tracked or exceeded the historical range between September and mid-November and set
new records in mid-November and mid-December 2020. On
21 November 2020, the maximum UVI measured on this day
exceeded the average of the daily maxima for 21 November,
calculated from measurements of the years 1991–2017, by
83%. At Arrival Heights, the UVI reached a new all-time site
record of 7.8 on 23 December 2020, exceeding the previous
record for this day by nearly 50%. Measurements at Palmer
Station were highly variable, as is typical for this site, but
new records were also set at this site in the second half of
November 2020 when the center of the ozone hole was above
the station. High UV radiation at this time, which coincides
with the start of the growing season for plants and the peak
breeding season for most animals, is a concern [3].
The record-high UVIs in 2020 were not only confined to
the three stations shown in Fig. 7 but also observed at other
Antarctic research stations. At the Australian Antarctic bases
Casey (66° S), Mawson (68° S), and Davis (69° S), UVIs
measured with broadband radiometers between October and
December 2020 were generally well above the 2007–2019
climatological mean, with new record-high values set on several days in November and December [31]. The number of
days when TCO dropped below 220 DU and led to spikes
in UVI was the highest ever observed at the three sites. The
daily maximum UVI at Marambio, a station located near the
Antarctic Peninsula at 64° S, exceeded 12 on several days
in late November and early December 2020 [20]. Similarly,
extreme UVI values were measured at King George Island
(62° S), near the northern tip of the Antarctic Peninsula [34].
959
The UVI exceeded 11 on four days between 24 November
and 4 December 2020 and peaked at 14.3 on 2 December.
This value ties, within the measurement uncertainty, with
the highest value of 14.2 (recorded on 4 December 1998)
ever measured at Palmer Station ([230] and Sect. 7.3). On
3 December 2020, the erythemal daily dose at King George
Island was 8.1 kJ/m2, which is among the highest on Earth
and only comparable to those recorded at high-altitude sites
such as the Atacama Desert, Chile [231], or at Mauna Loa,
Hawaii, where the highest dose ever observed was 9.5 kJ/m2
[232]. These extreme levels of UV radiation were a result of
solar elevations close to their annual maximum; close to 24 h
of daylight at King George Island; broken clouds, which can
enhance radiation levels at the surface beyond the clear-sky
level when the solar disk is free of clouds and additional radiation is scattered by clouds to the observer; and low TCO. For
example, on 1 December 2020 the TCO over King George
Island was 180 DU, which is the lowest value ever recorded
for December at this site [34]. UVI data for 2021 from stations
other than those shown in Fig. 7 are not yet available.
The findings of the studies discussed above show that
variability of springtime UV-B radiation in Antarctica is
large despite ongoing reduction of ODSs and signs of ozone
recovery. This surprisingly high variability is mainly driven
by changes in meteorological conditions and in particular
the persistently low temperature of the lower stratosphere.
When the Antarctic polar vortex breaks up at the end
of the austral spring, ozone-depleted air masses disperse to
lower latitudes, which may lead to large increases in UV
radiation over populated areas in the Southern Hemisphere
[233]. However, a recent study found that the breakup of the
polar vortex had only a small effect on UV radiation at Cape
Town, South Africa (34° S). Elevated levels of UV radiation
at this location were more frequently associated with lowozone air masses of tropical origin [234].
7.1.2 Temporal variations of UV radiation in the Arctic
As discussed in Sect. 3.3, an exceptionally large episode of
stratospheric ozone depletion was observed in late winter and
early spring (February–April) of 2020 in the Arctic [35, 36,
39, 41].
Figure 8a shows deviations of monthly average TCO from
past (2005–2019) averages north of 45° N for March, April,
May, and June, and their effects on UVI. In March 2020, relative TCO anomalies of up to − 40% and exceeding 3 standard
deviations (σ) were measured over northern Canada and the
adjacent Arctic Ocean. In April, relative TCO anomalies of
up to − 35% and exceeding 3σ were observed for virtually all
areas north of 60° N. During the breakup of the polar vortex in
May [35], areas with abnormally low (> 3σ) TCO still persisted
over Siberia.
13
960
Photochemical & Photobiological Sciences (2023) 22:937–989
Fig. 8 Monthly mean anomaly maps (in %) of a TCO and b noontime UVI for March, April, May, and June 2020 relative to 2005–
2019 means. Stippling indicates pixels where anomalies exceed three
standard deviations (3σ). Gray-shaded areas centered at the North
Pole in the maps for March and April indicate latitudes with no OMI
data because of polar darkness. Locations of ground stations are indicated by crosses in every map, with labels added to the first panel.
Maps are based on the OMTO3 Level 3 TCO product [237]. c Percentage anomalies in monthly means of the noontime UVI for 2020
derived from measurements at 10 ground stations (North to South
along the x-axis) relative to all years with available data (red) and
2005–2019 (blue). The black datasets indicate anomalies for the same
stations derived from OMI measurements relative to 2005–2019. Site
acronyms are ALT: Alert (83° N); EUR: Eureka (80° N); NYA: NyÅlesund (79° N); RES: Resolute (75° N); AND: Andøya (69° N);
SOD: Sodankylä (67° N); TRH: Trondheim (63° N); FIN: Finse
(61° N); OST: Østerås (60° N); and CHU: Churchill (59° N). Figure
adapted from Bernhard et al. [235]
The low TCO led to record-breaking anomalies in solar
UV-B radiation over the Arctic measured by ground-based
instruments at ten Arctic and subarctic locations and
observed by the Ozone Monitoring Instrument (OMI) on
NASA’s Aura satellite [235, 236]. Relative UV-B radiation
anomalies were particularly large between early March and
mid-April 2020. However, absolute anomalies for this period
remained small (e.g., below 0.6 UVI units) because solar
elevations for March and April are still low in the Arctic. In
the following, we only discuss relative anomalies.
In March 2020, the monthly average UVI over the Canadian Arctic and the adjacent Arctic Ocean was between 30
and 70% higher than the historical (2005–2019) averages,
often exceeding the climatological average by 3σ. By April
2020, they were positive over a vast area, including northern
Canada, Greenland, northern Europe, and Siberia. The maximum anomaly was 78% and anomalies exceeded 3σ almost
everywhere north of 70° N. In May 2020, UVI anomalies
of up to 60% and exceeding 3σ were measured over Siberia.
The UVIs in June were elevated by up to 30% over parts of
Norway, Sweden, and Finland, resulting from a combination
13
Photochemical & Photobiological Sciences (2023) 22:937–989
of negative TCO anomalies and unusually fair weather with
several cloudless days [236]. Ground-based measurements
generally confirm UVI anomalies derived from satellite data
(Fig. 8c). However, notable differences between the groundbased and satellite data sets exist for Sodankylä (Finland),
and Trondheim and Finse (Norway) in May. These discrepancies are likely caused by a mismatch between the albedo climatology used in the satellite retrieval and the actual albedo.
Albedo in May is affected by the timing of snow melting,
which was unusually late at Sodankylä and Finse in 2020.
In contrast to 2020, Arctic UVI anomalies in 2019 and
2021 remained within 2σ of the climatological mean, with
few exceptions [46, 47]. One exception is the large UVI
anomalies of up to 65% in the period 15–30 April 2019 in
Norway, Sweden, and Finland, when a persistent high-pressure system with clear skies was centered over the Nordic
countries. As the TCO in the Arctic is projected to have large
year-to-year variability for the remainder of the twenty-first
century, large variations in UV radiation are likely to occur
over the next decades.
7.1.3 Dependence of UV radiation on altitude
Measurements from satellites suggest that the highest UVI
values observed during the year at the Earth’s surface range
from less than 3 at the poles to about 25 at high altitudes
within the tropics of the Southern Hemisphere, such as
the Altiplano Region of Peru and Bolivia [210]. The average altitude of this region is 3750 m and the highest peak
(Illimani) is at 6438 m above sea level (asl). Ground-based
measurements of UV radiation in this area are sparse despite
their importance for human health and ecosystems. Recent
measurements at Quito, Ecuador (2850 m asl), established
a maximum UVI of 21 at this location [238]. This value is
consistent with the highest value of 21.2 measured at Mauna
Loa (3397 m asl) [232] and supports the maximum value
of ~ 25 for the highest UVI that may occur on Earth considering that Quito is at a considerably lower elevation than the
highest peaks of the Andes. The extreme UVI values at highaltitude locations close to the equator may have significant
health effects for people moving to these regions for work or
recreation without taking appropriate precautions to protect
themselves from UV radiation [239].
7.2 Observed long‑term changes in UV radiation
In the last four years, new trends in UV radiation derived
from ground-based measurements have been published
for several regions [190, 223, 238, 240–245]. These studies confirm that changes in UV radiation during the last
25 years have generally been small—typically less than 4%
per decade, increasing at some sites and decreasing at others,
with few exceptions—consistent with the multi-site study
961
by McKenzie et al. [56] discussed in Sec 4.1. Results from
these studies are assessed in more detail below. While only
studies that appeared to be of high quality according to our
assessment were included, the measurement uncertainty of
the various datasets varies and the reader is referred to the
original publications for details.
Trends in solar spectral irradiance at 307.5 nm, which
is a reasonable proxy for trends in erythemally weighted
UV radiation, were calculated at several stations in Europe,
Canada, and Japan over a 25-year (1992–2016) period [240].
Long-term changes at this wavelength vary by location and
are mostly driven by changes in aerosols and TCO. However,
at high northern latitudes, changes in the surface reflectivity
are also an important factor. Over Japan, the spectral irradiance at 307.5 nm has increased significantly by about 3% per
decade over this 25-year period and this increase is attributed to a decrease in absorbing aerosols. The only European
station with a significant trend was Thessaloniki, Greece,
where spectral irradiance at 307.5 nm rose by 3.5% per decade with an increasing rate of change during the last decade,
possibly because of decreasing absorption by aerosols.
Updated estimates of trends in UV-B irradiance at
four European stations (Reading (51° N), Uccle (51° N),
Sodankylä (67° N), and Thessaloniki (41° N)) have been
reported for the period 1996–2017, i.e., starting after the
global peak of ODSs [245]. The study concluded that the
variability of UV-B radiation at these European sites was
mainly governed by variations in clouds, aerosols, and surface reflectivity, while changes in TCO were less important.
Statistically significant (95% confidence level (CL)) positive trends in noontime spectral irradiance at 307.5 nm were
found for Thessaloniki (8% per decade) and Uccle (5% per
decade), while, for Reading, the trend was negative (− 7%
per decade). These trends were again attributed to the effects
of aerosols and clouds. No statistically significant trend
was found at Sodankylä; however, the decreasing tendency
of − 5% per decade at this site was found to be consistent
with changes in surface reflectivity due to declining snow
cover in late winter and spring. In a follow-on study [223],
a similar trend analysis was performed for Rome (42° N),
Italy. A statistically significant negative trend in TCO of
–1% per decade was found, but there was no corresponding significant increase in spectral irradiance at 307.5 nm
over the period 1996–2020. However for certain months,
positive trends in UV irradiance were observed, which were
predominantly caused by changes in clouds and/or aerosols.
Several other studies reported estimates of trends for erythemal irradiance (or erythemal doses) at northern European
sites. No statistically significant trends in erythemal UV
radiation were observed at Moscow, Russia (56° N), over
the period from 1999 to 2015 [243]; at Chilton, England
(52° N), between 1991 and 2015 [246]; and at Tõravere,
Estonia (58° N), between 2004 and 2016 [244]. At the last
13
962
site, there were also no trends in the main factors influencing UV radiation, namely TCO; aerosol optical depth; and
global short-wave radiation, which is a proxy for the effect
of clouds.
Trends in erythemal irradiance at the Earth’s surface over
the period of 2005–2017 have been calculated for the continental United States using satellite-based (OMI) measurements and ground-based measurements at 31 sites distributed
throughout the United States by the Department of Agriculture’s UV-B Monitoring and Research Program [241]. The
study concluded that trends in noontime erythemal irradiance
estimated from these satellite- and ground-based measurements cannot be reconciled. Specifically, trends derived from
the satellite-based dataset were not significant for most of the
continental United States, except for a small region in the
New England states of Maine, Vermont, New Hampshire,
and Massachusetts. In those regions, small (about 5% per
decade) positive trends were calculated from OMI data, and
they were significant at the 95% CL. However, data from the
two ground-based stations located in this region indicated a
significant decrease in erythemal UV over the same period.
This discrepancy can be explained, either by calibration
issues of the ground-based sensors and OMI [247], or by
increasing attenuation of UV radiation in the lowest part of
the atmosphere, which cannot be adequately probed by OMI.
While trends calculated for several other stations were also
significant, the magnitude of these trends is generally within
the measurement uncertainty range so that no firm conclusions about changes in levels of erythemal irradiance across
the continental United States can be drawn.
In a similar study based on OMI measurements, trends in
noontime erythemal irradiance, TCO, and cloud and haze
transmission were calculated for 191 cities located between
latitudes of 60° N and 60° S over the period 2005 − 2018
[248]. Significant changes in erythemal irradiance were
found at the 95% CL for 40 of the 191 sites over this period.
When data were averaged over 15° latitude bands, correlations between erythemal irradiance and short- and longterm changes in cloud and absorbing aerosols, as well as
inverse correlations between UV radiation and TCO, became
apparent. Estimates of changes in atmospheric transmission at 340 nm show increases of 1.1 ± 1.2% per decade
between 60° S and 45° S, almost no change between 45°
S and 20° N, decreases of 3% per decade at 22° and 32° N,
an increase of 2.5% per decade near 25° N, and increases
of 1 ± 0.9% per decade from 35° N to 60° N. Changes in
zonally averaged (~ 15° latitude bins) erythemal irradiance
between 60° N and 60° S range between –4 and 5% per decade and are predominantly caused by changes in cloud and
aerosol transmission. However, judging from the error bars
in the figures provided by Herman et al. [248], changes in
zonally averaged transmission and erythemal irradiance are
generally not significant at the 95% CL.
13
Photochemical & Photobiological Sciences (2023) 22:937–989
Trends in erythemal daily doses (Dery) were calculated
for the period 1979–2015 over Northern Eurasia (a region
between 40° and 80° N and extending in longitude from
Scandinavia to Siberia) using simulations by a climate chemistry model (INM-RSHU34), re-analyses of atmospheric data
(ERA-Interim35), and data from satellite measurements
(TOMS/OMI) [242]. For cloud-free conditions, statistically
significant increases in Dery of up to 3% per decade were
found for spring and summer over large areas and attributed to decreases in TCO. When clouds were included in the
analysis, greater trends of 6–8% per decade were found over
Eastern Europe and several regions in Siberia and Northeast Asia. This observation suggests that over this 36-year
period, changes in cloud attenuation had larger effects on
UV-B radiation than changes in TCO at high latitudes of the
Northern Hemisphere extending to 80° N.
An analysis of UVI data computed from satellite-based
measurements for local noon and clear skies by the Tropospheric Emission Monitoring Internet Service (TEMIS) indicated that there is no long-term trend in UVI at the equatorial high-altitude site of Quito, Ecuador (0° S, 2850 m
asl), for the period 1979–2018 [238]. This conclusion was
corroborated by ground-based measurements at this site.
For 2010–2014, the measured UVI was within the range of
variability inferred for 1979–2009 from TEMIS data. This is
consistent with the observation that there are no significant
trends in TCO in the tropics (Sect. 3.1).
Trends in the UVI measured by spectroradiometers at
three Antarctic sites (South Pole (90° S), Arrival Heights
(78°S), and Palmer Station (65°S)) have recently been reassessed for the period of 1996–2018 [190]. At the South Pole
(a site representative of the Antarctic Polar Plateau), significant (95% CL) decadal trends of − 3.9% and − 3.1% were
calculated for January and February, respectively, which
can mostly be explained by concomitant trends in TCO. At
Arrival Heights, the recalculated trend for summer is − 3.3%
per decade and is significant at the 90% CL. This downward
trend is caused by a significant upward trend in TCO of 1.5%
per decade for January plus the effect of reductions in landfast ice covering the sea adjacent to the instrument site (Sec
6.2). No significant trends were reported for Palmer Station.
The study provides further evidence that the UVI in Antarctica is starting to decrease during summer months. However,
statistically significant reductions for spring (October and
34
INM-RSHU: Institute of Numerical Mathematics – Russian State
Hydrometeorological University; ERA-Interim: European Centre for
Medium-Range Weather Forecasts Re-Analysis; TOMS: Total Ozone
Mapping Spectrometer.
35
ERA-Interim is a global atmospheric reanalysis published by the
European Centre for Medium-Range Weather Forecasts (ECMWF,
https://www.ecmwf.int/) that is available from 1 January 1979 to 31
August 2019.
Photochemical & Photobiological Sciences (2023) 22:937–989
7.3 Reconstruction of historical changes in UV
radiation
Systematic measurements of surface UV radiation suitable for trend analysis began only in the late 1980s. In the
absence of direct measurements, knowledge of UV irradiance levels prior to the onset of ozone depletion relies on
radiative transfer model calculations in combination with
inputs such as TCO and other proxy data. At very few locations, ground-based TCO measurements commenced before
the 1960s and UV irradiances have been reconstructed from
these measurements [249–252]. The erythemal UV irradiance was recently reconstructed for Moscow, Russia, for the
warm season (May–September) over the period 1968–2016
[253] using data of TCO, AOD at 550 nm, surface albedo,
cloud cover, and cloud transmission. Results were validated
against measurements of broadband instruments emulating
the erythemal response of human skin (Sect. 10.1), which
were available from 1999 onward. Reconstructed and measured data for the overlap period agreed well; the coefficient
of determination R2 was 0.89. Results indicate statistically
significant decadal trends in erythemal UV irradiance of
–11.6 ± 1.6% for the period 1968–1978 and 5.1 ± 1.9 for
the period 1979–2016, which were predominantly driven
by changes in cloud transmission. One important shortcoming of the study is that the consistency of cloud data of this
48-year data record was not independently verified; hence,
trend estimates could be affected by spurious trends in the
measures of cloudiness.
Daily erythemal UV doses were reconstructed for Novi
Sad, Serbia [254]. Using a radiative transfer model with
inputs of TCO and snow cover data, plus empirical relations
between erythemal doses and sunshine duration, statistically
significant increases in erythemal UV doses of 8.8% and
13.1% per decade over the period 1980–1997 were found
for summer and winter, respectively, which were linked to
the statistically significant decline in TCO over this period.
Satellite measurements of TCO became available in
the late 1970s and have also been used for reconstructing
the UVI at several ground stations under the assumption
that changes in aerosol and clouds were small during this
Seasonal Changes in the UV Index
Northern Hemisphere month
J
16
F
M
Winter
A
M
J
Spring
J
A
S
Summer
O
N
D
Autumn
14
12
Maximum UV Index
November), when the ozone hole leads to large UVI variability, were not detected.
All studies summarized above paint a consistent picture:
changes in UV-B radiation outside the polar regions over
the last 2–3 decades are mainly governed by variations
in clouds, aerosols, and surface reflectivity (for snow- or
ice-covered areas), while changes in TCO are less important. These results corroborate the conclusion by McKenzie
et al. [56] discussed in Sect. 4.1 that changes in TCO have
not led to significant changes in UV-B radiation over this
period.
963
10
8
6
4
2
0
J
A
S
O
N
D
J
F
M
A
M
J
Southern Hemisphere month
Maximum UV Index on a given day of the year
.
Pre-ozone hole
(calculated)
Since early 1990s.
(measured)
Palmer, Antarctica (64 S)
1970 – 1976
1990 – 2020
San Diego, California (32 N)
1970 – 1976
1992 – 2008
Utqiagvik, Alaska (71 N)
(formerly Barrow)
1970 – 1976
1991 – 2016
Fig. 9 Comparison of the highest UVIs ever measured for each
day of the year at Palmer Station, San Diego, and Barrow since the
early 1990s (solid lines) with reconstructed data for the pre-ozonehole period 1970–1976 (broken lines). Yellow shading indicates the
change between historical and contemporary UVI. The difference is
particularly large for Palmer Station during spring, the period affected
by the Antarctic ozone hole. The highest UVIs observed at Palmer
since the 1990s exceed those measured at San Diego despite that
city’s much lower latitude. Reprinted from Bernhard et al. [230]
period [56]. These reconstructions imply that considerable
increases in the summer UVI occurred between 1978 and
1990, ranging from about 5% at northern mid-latitudes, up
to 10% at southern mid-latitudes, and up to 20% at the three
Antarctic sites considered in this study.
Starting in 2010, the “Twenty Questions and Answers
About the Ozone Layer” component of assessment reports
prepared by the SAP have included a plot comparing reconstructed UVIs at Palmer Station, Antarctica (64° S), for the
pre-ozone-hole period 1978–1980 with UVIs measured
between 1990 and 2006 [13, 255, 256]. This plot has recently
been updated [230] and is reproduced in Fig. 9. The revised
plot is similar to the legacy one but includes data up to 2020
and also compares recent measurements with reconstructed
pre-ozone-hole UVIs for San Diego, California (32° N),
and Barrow, Alaska (79° N). Furthermore, historical UVIs
at the three sites have been calculated from TCO measurements by the Backscatter Ultraviolet (BUV) experiment
13
964
on the Nimbus-4 satellite between 1970 and 1976. While
trends in TCO were already negative in the 1970s over polar
regions [67], analysis presented by Bernhard et al. [230]
did not show clear evidence that the developing ozone hole
affected Palmer Station before 1976. In contrast, the period
1978–1980 used for the legacy plot was already somewhat
influenced by ozone depletion. The new results confirm the
previous conclusion that the ozone hole led to large increases
in the UVI at Palmer Station year-round, with the largest
increases occurring during spring (between 15 September
and 15 November). The maximum UVI at this site is now
larger by a factor of 2.50 ± 0.37 (± 1σ) on average compared
to the pre-ozone-hole period. During summer and autumn
(21 December–21 June), i.e., the seasons least affected by
the ozone hole, UVI maxima measured between 1990 and
2020 exceed maxima estimated for years prior to 1976 by
20 ± 13%. Measured and reconstructed pre-ozone depletion
data for San Diego (a subtropical site), are almost indistinguishable: on average, the UVI has increased by 3 ± 7%
(± 1σ) since the 1970s. This modest growth is consistent with
the small change in TCO observed at subtropical latitudes
(Sect. 3.1) and with the conclusion of McKenzie et al. [56]
that maximum daily UVI values have remained essentially
constant at mid-latitudes over the last ~ 20 years due to the
phase-out of ODSs controlled by the Montreal Protocol. At
the Arctic site of Barrow, the UVI increased by 18 ± 15%
(± 1σ) since the 1970s. The largest spikes in the UVI of up
to 40% relative to the 1970s were measured during spring in
years with abnormally strong Arctic ozone depletion, such
as 2011 [257]. We note that these reconstructions are subject
to uncertainty because they assume that surface albedo and
attenuation by clouds and aerosols have not changed over the
last 50 years in this area. However, at Palmer and Barrow,
the TCO is by far the most important factor in controlling
the UVI, while changes in albedo at San Diego can be considered negligible. Note that changes in the UVI discussed
here do not contradict the conclusion in Sect. 7.2 that longterm changes in UV-B radiation outside the polar regions
have generally been small over the last 2–3 decades. Changes
shown in Fig. 9 are by and large attributable to changes in
TCO occurring in the 1970s and 1980s (Fig. 1).
At Athens, Greece, records of the duration of sunshine
were used to reconstruct monthly averages of short-wave
(wavelength range ~ 300–3000 nm) solar irradiance at the
surface between 1900 and 2012 [258]. There were very
small (0.02%) changes between 1900 and 1953, followed
by a negative trend of 2% per decade during a “dimming”
period of 1955–1980 and a positive trend of 1.5% per decade
during a “brightening” period of 1980–2012. Measurements
of short-wave irradiance at Potsdam (52°), Germany, show
distinct dimming and brightening periods between 1947 and
1986 and 1986–2016, respectively, with measurements in
1986 about 10 W/m2 lower compared to those at the start
13
Photochemical & Photobiological Sciences (2023) 22:937–989
and end of the time series [259]. Changes for “all-sky”
(cloudy and cloud-free) and clear-sky (cloud-free) conditions were similar, suggesting that changes in aerosols were
mostly responsible for these variations in short-wave irradiance. While these trend estimates are unrelated to changes
in TCO and do not directly translate to changes in UV
radiation, they qualitatively capture variations in the effect
of clouds and aerosols on solar irradiance over the period
studied, which are also relevant for changes in UV radiation.
Trends in UV radiation related to changes in aerosols are
likely larger than trends in short-wave irradiance because
attenuation by aerosols is generally larger in the UV region
than at visible wavelengths.
8 Projections of UV radiation
Projections of solar UV radiation at the Earth’s surface for
the twenty-first century have been reported in new studies
published during the last four years. These new projections
are generally in agreement with those reported in our last
assessment [9]. They confirm the projected reductions in
UV radiation, particularly at high and polar latitudes, due to
the recovery of stratospheric ozone, as well as the increases
in UV radiation due to decreasing concentrations of aerosols over regions with intense urban or industrial activities.
Furthermore, projected decreases in surface reflectivity due
to reduction in ice cover and decreases in cloudiness, both
associated with climate change, are also important drivers
leading to regional changes (decreases and increases, respectively) in surface UV radiation.
One of the new studies [260] reports global projections
of UVI that were calculated with a radiative transfer model
using TCO, temperature, and aerosol fields provided by 17
CCMs. These CCMs were included in the first phase of the
Chemistry-Climate Model Initiative (CCMI-1) [58, 261].
The CCM simulations were performed for four future GHG
scenarios described by RCPs 2.6, 4.5, 6.0 and 8.5. Zonalmean noontime UVI for cloudless skies were calculated for
the period 1960–2100.
According to this study, noontime UVI in 2100 is projected to increase relative to calculations for the 1960s for
RCPs 2.6, 4.5 and 6.0. These increases depend on latitude
and the RCP scenario, and range between 1% (northern high
latitudes for RCP 6.0) and 8% (northern mid-latitudes for
RCP 2.6) as shown in Table 1. Trends calculated for the
worst-case scenario RCP 8.5 show a different pattern with
UVI projected to increase only in the tropics and to decrease
elsewhere, with the largest decrease of 8% at northern high
latitudes.
Only three of the 17 CCMs provided outputs of the
AOD and its wavelength-dependence. The AOD used
Latitudes
Lamy et al. [260]
Bais et al. [9]
a
b
Change [%] 2015 to 2090
Change [%] 2015 to 2090c
Fixed AODs
Transient
AODs
Fixed AODs
Fixed AODs
RCP 6.0
RCP 6.0
RCP 6.0
RCP 6.0
Change [%] 1960 to 2100
Transient AODs, RCP =
2.6
4.5
6.0
8.5
e
Annual mean
High Nd
30–60° N
0–30° Nf
0–30° Sf
30–60° S
High Sd
6
8
3
3
3
7
2
5
3
3
3
6
1
5
3
3
2
4
−8
–1
1
2
–2
0
–5
–2
3
3
0
–2
0
5
–
0
–5
– 18
–6
–3
1
0
–6
– 18
Jan
Apr
Jul
Oct
–3
–4
–1
–1
–5
–8
–7
–5
0
0
–4
–6
–5
–3
–1
–1
–5
–6
–4
–2
–1
–1
–6
– 23
Photochemical & Photobiological Sciences (2023) 22:937–989
Table 1 Comparison of zonal mean changes in clear-sky UVI calculated by Lamy et al. [260] and Bais et al. [9]
All values are rounded
a
According to Table 5 of Lamy et al. [260]
b
Inferred from Figs. 4, 6 of Lamy et al. [260]
c
Table 1 of Bais et al. [9]
d
Latitude range of “High N” and “High S” refer to 60–90° in Lamy et al. [260] and 60–80° in Bais et al. [9]
e
Changes reported by Lamy et al. [260] refer to trends averaged over all months; Bais et al. [9] provide changes for the months of January, April, July, and October
f
Lamy et al. [260] report changes separately for 0–30° N and 0–30° S while Bais et al. [9] provide changes for 30° N to 30° S
965
13
966
in projections of UV radiation is therefore based on the
median of AODs derived from these three CCMs. According to these calculations, AODs are projected to decrease by
almost 80% between 2000 and 2100 at northern high- and
mid-latitudes, resulting in concomitant increases in the UVI
of about 2% and 6%, respectively. These changes in UVI due
to changes in AOD are of similar magnitude to those caused
by changes in stratospheric ozone. However, these AOD estimates as well as the absorption properties of aerosols used in
these CCMs are highly uncertain because future changes in
atmospheric aerosols depend greatly on policy choices, such
as measures to reduce air pollution [262]. Moreover, changes
in optical depth and absorption properties of aerosols are
highly dependent on region, hence zonal mean changes in
UVI, like those discussed above, are not necessarily representative for most regions.
To address these concerns, Lamy et al. [260] also provide
UVI projections for temporally invariant or “fixed” AODs
based on a current climatology [263]. Using this climatology
and the RCP 6.0 scenario, noontime UVIs in 2100 are projected to change relative to 1960 by –5% at Northern Hemisphere (NH) high latitudes, –2% at NH mid-latitudes, + 3%
in the tropical belt, 0% at Southern Hemisphere (SH) midlatitudes, and –2% at SH high latitudes (Table 1, column 6).
These changes in UVI are mainly driven by changes in TCO.
Assuming time-invariant aerosol amounts for the future, the
clear-sky UVI is projected to decrease from 2015 to 2090 by
3% at NH and 6% at SH mid-latitudes. However, in regions
that are currently affected by air pollution, the UVI is projected to increase if emissions of air pollutants are curtailed
in the future.
Table 1 also shows a comparison of projected changes
in zonal mean UVIs between 2015 and 2090 inferred from
the study by Lamy et al. [260] and as published in our last
assessment [9]. In both cases, UVI projections were based on
results from the CCMI-1 initiative; however, different subsets
of models were used, as well as methods to calculate the
UVI from parameters provided by the CCMs. Furthermore,
Bais et al. [9] provided projections for different months while
Lamy et al. [260] only considered annual averages. Despite
these differences, changes in UVI calculated by the two studies for fixed AODs are consistent (see last five columns of
Table 1) and project a decrease of 2–5% for northern midlatitudes, a decrease of 4–6% for southern mid-latitudes, and
almost no change for the tropics. Both studies also predict
large decreases in the UVI over southern high latitudes due
to the expected healing of the stratospheric ozone hole.
Projections provided in the above studies were corroborated by another study where long-term changes in erythemal UV radiation were calculated over Eurasia (latitudes
40–80° N, longitudes 10° W–180° E) based on results of a
CCM developed by the Russian State Hydrometeorological
University (RSHU) [264]. These calculations considered
13
Photochemical & Photobiological Sciences (2023) 22:937–989
only changes in TCO (i.e., excluding effects of aerosols and
clouds) and predict that erythemal UV radiation levels in
the years 2055–2059 will be lower over Eurasia by 4 to 8%
relative to the reference period 1979–1983.
Simulations with one of the CCMs (EMAC 36) for the
period 1960–2100 were used to derive trends in DNA-damaging radiation at four mid-latitude locations and one tropical high-altitude site [265]. Weighting the spectral irradiance
with the action spectrum for DNA-damage [266] yields dose
rates that are more sensitive to changes in radiation at shorter
(UV-B) wavelengths than the erythemal UV dose rates or the
UVI; hence it is more sensitive to changes in TCO. DNAdamaging irradiance averaged over the five locations considered in this study is projected to increase by 1.3% per decade
between 2050 and 2100. To isolate the effect of GHGs on
climate, one simulation assumed increasing GHGs according to RCP 6.0 and the second adopted constant GHGs at
1960 levels. No trend in TCO was detected by the model
after 2050, and the trend detected in DNA-damaging irradiance was attributed to a statistically significant (95% CL)
decrease in cloud cover of 1.4% per decade resulting from
increasing GHGs. The study suggests that changes in UV-B
irradiance at low- and mid-latitudes during the second half
of the twenty-first century will be dominated by factors other
than changes in stratospheric ozone. However, these projections depend on the accuracy of simulating the cloud fields
by climate models because uncertainties in the modeling
of clouds propagate to the projected changes in solar UV-B
radiation.
The SAP’s latest assessment [11] also evaluates the effect
of a fleet of commercial supersonic aircraft on stratospheric
ozone concentrations. Such a fleet is currently being considered by different organizations and companies. Depending
on scenario and flight altitudes, emissions of water vapor
and nitrogen oxides from such a fleet could reduce TCO by
up to 25 DU at high northern latitudes. Reductions in TCO
at mid and low latitudes of the Northern Hemisphere would
be considerably smaller, and the Southern Hemisphere is
less affected because most flights take place in the Northern
Hemisphere. While no study has quantified the effect of a
future fleet of supersonic aircraft on UV radiation, the estimated decrease in TCO suggests that erythemal UV irradiance could increase by several percent at mid-latitudes of the
Northern Hemisphere.
New model calculations examined the effect on stratospheric ozone (and by implication on UV radiation) of quadrupling concentrations of atmospheric CO2 [267]. Such an
increase would lead to a dynamically-driven decrease in
concentrations of ozone in the tropical lower stratosphere,
36
EMAC CCM is the European Centre for Medium-Range Weather
Forecasts–Hamburg (ECHAM)/Modular Earth Submodel System
(MESSy) Atmospheric Chemistry Model.
Photochemical & Photobiological Sciences (2023) 22:937–989
an increase of ozone in the lower stratosphere over the high
latitudes, and a chemically driven increase of ozone (via
stratospheric cooling) throughout the upper stratosphere. In
the tropics, opposite changes in ozone in the upper and lower
stratosphere result in small changes in the TCO, and, in turn,
to small changes in tropical UV-B radiation in the future, if
effects from all other factors remain the same. A quadrupling
of atmospheric CO2 concentrations during the twenty-first
century is currently not expected, but could occur in the
twenty-second century if emissions of CO2 were to continue unabated according to the RCP 8.5 scenario [268].
The study suggests that even “worst-case” increases in CO2
will not result in significant increases in UV-B radiation in
the tropics.
All studies discussed above confirm the understanding of
UV radiation in the twenty-first century established in our
last assessment [9]. Simulations with a new generation of
CCMs that have only recently been performed are expected
to provide updated projections of UV radiation but are not
yet available for assessment in this report.
A recent study used a state-of-the-art climate model with
interactive chemistry [269] to calculate the effects on TCO
and UV radiation resulting from a regional or global nuclear
war. A global-scale nuclear war would cause a 15-year-long
reduction in the TCO with a peak loss of 75% globally
and 65% in the tropics. Initially, soot would shield the surface from UV-B radiation, but eventually the UVI would
become extreme: greater than 35 in the tropics for 4 years,
and greater than 45 during the summer in the southern polar
regions for 3 years. For a regional nuclear war, global TCO
could be reduced by 25% with recovery taking 12 years.
9 Implications of solar radiation
management on UV radiation
Over the last decade, global warming from increasing GHGs
has accelerated, and global mean air temperatures near the
surface have risen by about 1.1 °C above pre-industrial
levels [Chapter 2 of 63]. The resulting changes in climate
observed worldwide have stimulated discussions on strategies to mitigate warming through artificially forced reduction of solar radiation entering the troposphere. Impacts of
such solar radiation management (SRM) interventions on the
atmosphere and the environment have been investigated in
numerous modeling studies and discussed in current assessments by the SAP [11] and IPCC [Chapter 4 of 63], and
the last EEAP assessment [9]. The latest SAP report [11]
extensively addresses the potential impacts on TCO from
stratospheric aerosol injection (SAI) under different scenarios. Here, we focus on the effects of the possible implementation of SAI on surface UV radiation. The effects are driven
not only by changes in TCO but also by the redistribution of
967
solar radiation from the direct-to-diffuse component, plus
the global dimming effect expected from back-scattering of
solar radiation to space by the aerosol layer.
The TCO is affected both by SAI-induced changes in heterogeneous chemical reactions, which depend on the surface
area density of the aerosol (e.g., in µm2/cm3), and by changes
in atmospheric dynamics (including transport, temperature,
and water vapor changes). These effects on TCO differ with
latitude and season, and depend on the future SAI scenario
because they act in addition to the effects of decreasing
ODSs and increasing GHGs. During the Antarctic ozone
hole season, destruction of ozone in the stratosphere resulting from SAI would mainly be controlled by halogen chemistry on the surface of aerosols, while transport of ozone
through circulation becomes important in other seasons [11].
Using models that participated in the Geoengineering
Large ENSemble (GLENS) project, Tilmes et al. [270] estimated the effect on TCO in the latitude band 63°–90° S
from SAI designed to achieve a reduction of 1.5 and 2.0 °C
in global surface temperature. They found a reduction of
up to 70 DU in the Antarctic TCO at the start of the SAI
application (2020–2030), followed by an increase in TCO
towards 2100 with a pattern like the projected changes in
TCO without the application of SAI. In a more recent study,
Tilmes et al. [271] estimated the initial abrupt decrease in
TCO to be between 8 and 20% in 2030–2039 compared to
2010–2019, depending on injection strategy and model.
All scenarios assumed in these studies result in a delayed
recovery of Antarctic ozone to pre-ozone-hole levels by
20 to ~ 40 years. The TCO for these SAI scenarios remains
below the levels projected by the worst case GHG scenario
(SSP5-8.5) until the end of the twenty-first century, which
would lead to increased levels of UV-B radiation during the
entire period in Antarctica.
In a similar study, Tilmes et al. [272] estimated the
effects of SAI also in the Northern Hemisphere and the
tropics based on simulations of the G6 Geoengineering
Model Intercomparison Project (GeoMIP). The models
agree that sulfur injections result in a robust increase in
TCO in winter at middle and high latitudes of the Northern
Hemisphere of up to 20 DU over the twenty-first century
compared to simulations based on the SSP5-8.5 scenario
without SAI. This increase in TCO, which is linearly related
to the increase in the amount of sulfur injections, is driven
by the warming of the tropical lower stratosphere and would
eventually result in decreasing UV-B radiation at these latitudes during the remainder of the twenty-first century. The
magnitude of these changes in UV-B radiation depends on
the SAI scenario. The Arctic TCO is initially projected to
decrease by 13 to 22 DU depending on the scenario, which is
a much smaller decrease than that projected by Tilmes et al.
[270] for the Antarctic discussed above. By the end of the
twenty-first century, the Arctic TCO with and without SAI
13
968
are approximately the same. Finally for the tropics, changes
in ozone due to SAI would be small. The initial reduction
in TCO projected by Tilmes et al. [270] and Tilmes et al.
[272] for the Antarctic and Arctic is attributable to heterogeneous reactions on aerosol particles in the presence of
ODSs. Robrecht et al. [273] showed that this effect is far
less important for mid-latitudes and the tropics compared
with polar regions.
While the above studies have focused on the consequences of SAI on ozone, effects on UV and visible radiation from SAI also depend on the attenuation (dimming)
and redistribution of solar radiation. These effects have been
quantified with a radiative transfer model using inputs from
the GLENS project [271] designed to counteract warming
from increased GHGs under the RCP 8.5 scenario [274].
Estimated changes in the UVI are predominantly driven by
the attenuation of solar radiation by the artificial aerosol
layer (with concentrations peaking above ~ 30 km in the tropics and above ~ 25 km in the high latitudes). Reduced direct
radiation due to aerosol scattering results in substantial
reductions in solar irradiance at the Earth’s surface despite
an enhanced contribution from diffuse radiation. However,
the larger diffuse component may allow more efficient penetration of UV irradiance through forest and crop canopies
[275], offsetting, to some extent, the reduced irradiance on
top of the canopies. The intervention is estimated to reduce
the daily average above-canopy UVI in 2080 relative to 2020
by about 15% at 30° N and by 6–22% at 70° N, depending on
season. About one third of the reduced UVI at 30° N is due
to the relative increase in TCO (~ 3.5%) between the reference and the SRM scenario. The corresponding increase in
TCO for 70° N is less than 1% and explains only a very small
fraction of the decrease in the UVI. The calculated changes
in the UVI are therefore primarily caused by the scattering effect of sulfate aerosols, with a very small contribution
from the absorption by sulfur dioxide (SO2). Finally, reductions in photosynthetically active radiation (PAR) are estimated to range from 9 to 16% at 30° N and from 20 to 72%
at 70° N, depending on season, with the largest proportional
changes occurring in December, when the absolute levels of
radiation are small. Such large changes in the UVI and PAR
would likely have important consequences for ecosystem
services and food security; however, such repercussions have
not yet been quantified. While the study only characterized
changes in UV radiation and PAR for the NH, similar results
can be expected for the SH.
13
Photochemical & Photobiological Sciences (2023) 22:937–989
10 Advances in UV monitoring
and modeling
In this section, we provide a summary of advances in measuring and modeling UV radiation at the Earth’s surface
and in assessing personal exposure to UV radiation, which
is controlled both by ambient UV radiation and personal
behavior.
10.1 Ground‑based systems
UV radiation at the Earth’s surface is normally measured
with scanning spectroradiometers, such as those installed
in the Network for the Detection Atmospheric Composition
Change (NDACC) [276]; broadband instruments, which
typically emulate the erythemal response of human skin
[277]; multi-filter instruments, which measure the spectral irradiance at several wavelengths (typically 4–7) in the
UV range [278]; array spectroradiometers, which record
the entire UV spectrum within seconds; dosimeters, which
measure the UV dose that accumulates over a given amount
of time; and specialized systems designed for a specific
research question such as the measurement of the angular
distribution of sky radiance [279]. The different instruments
have been discussed in detail in previous assessments [9,
162]. In brief, scanning spectroradiometers using double
monochromators are the most accurate instruments but are
expensive to acquire and maintain, and the recording of a
UV spectrum may take several minutes. Broadband radiometers are relatively inexpensive, and their spectral response
is tailored to a specific effect (e.g., erythema) under study,
but because they do not provide spectral information, the
factors driving changes in UV radiation (e.g., ozone, clouds,
and aerosols) cannot be unambiguously separated. Multifilter instruments can be used for studying a specific effect
and the factors it depends upon, but require elaborate characterisations and calibrations to provide accurate data of
solar irradiance [280]. Array spectroradiometers (or spectrographs) use single monochromators for physical reasons,
and measurements at wavelengths shorter than 310 nm are
often affected by stray light [281]. An instrument combining
an array spectrometer with narrow-band filters that mitigate
this problem has recently been introduced [282] and evaluated by [283], indicating good performance at wavelengths
longer than 305 nm. Finally, dosimeters are simple, lowcost, small devices that measure the UV dose electronically
[284], chemically [285, 286], or both [287], and are further
discussed in Sect. 10.5.2. Their accuracy is typically less
than that of high-end spectrometers [288]; however, they
are frequently used for exposure studies (Sect. 10.5.2) where
Photochemical & Photobiological Sciences (2023) 22:937–989
they can be easily attached, for example, to the forehead,
wrist or clothing of test subjects.
The quality of measurements of UV radiation from these
systems or sensors has historically been assessed with intercomparison campaigns where instruments are either compared with each other or a reference instrument. An example
of the latter is a campaign with 75 participating broadband
radiometers with erythemal response [277]. The instruments’
solar measurements were first compared with data from the
QASUME (Quality Assurance of Spectral Ultraviolet Measurements in Europe) reference spectroradiometer [289]. The
QASUME instrument has been used since 2002 to assess
the quality of UV radiation measurements from more than
250 spectroradiometers at more than 40 stations worldwide
(https://www.pmodwrc.ch/en/world-radiation-center-2/wccuv/qasume-site-audits/). New calibrations were subsequently
transferred from QASUME to the 75 broadband radiometers.
Furthermore, the angular and spectral response of the instruments was measured and functions for correcting deviations
from the ideal response were established. With their original calibration applied, measurements of 32 (43%) of the
75 instruments agreed to within ± 5% with measurements
of the reference spectroradiometer while 48 (64%) agreed
to within ± 10%. Twenty-seven (35%) datasets deviated by
more than ± 10% from the reference and two datasets differed by 70%. After instruments were recalibrated, 73 (97%)
of the 75 instruments agreed to within ± 5% with the reference. This example demonstrates that proper quality control,
quality assurance, and calibration procedures are vital for
obtaining accurate measurements of UV radiation. A similar
intercomparison involving four broadband radiometers and a
reference spectroradiometer was conducted between March
2018 and February 2019 at Saint-Denis, La Réunion (21°
S) [290]. Data from three of the four instruments agreed to
within ± 3% with the reference while data from one instrument exhibited a systematic error of 14%.
Even high-end spectroradiometers require meticulous
characterization and calibration for obtaining measurements with low uncertainty [291]. Finally, the development
of a rigorous uncertainty budget (i.e., the calculation, tallying and combination of all uncertainty components) is a
demanding task [292], but is necessary for obtaining high
quality data.
10.2 Modeling of UV radiation
The transfer of radiation through the Earth’s atmosphere is
affected by absorption and scattering by gases, aerosols, and
clouds; the reflection of radiation by the Earth’s surface; and
several other factors (Sect. 6). These factors are taken into
account in computer simulations of UV radiation by radiative transfer models. Physically correct radiative transfer
codes for modeling the UV radiation at the Earth’s surface
969
have been available for many years [e.g., 293,294–297]
and can be considered reliable and mature. Most models
assume that the atmosphere is homogeneous in both horizontal directions and only varies in the vertical direction, but
newer models (e.g., [298, 299]) that are based on the Monte
Carlo technique [300] can also account for the three-dimensional structure of the atmosphere, topography, surface condition (e.g., patchy snow) or illumination geometry (e.g.,
the inhomogeneous irradiation during a solar eclipse). The
greatest challenge in radiative transfer calculations is not the
physical description of the transfer of radiation through the
atmosphere but the specification of the input parameters that
interact with radiation and are often not completely known,
such as the single scattering albedo (SSA) of aerosols in the
UV-B range or the structure of clouds.
One source of uncertainty in determining the UV radiation at the Earth’s surface with models is the uncertainty
of the solar spectrum outside the Earth’s atmosphere. The
extraterrestrial solar spectra (ETS) used in legacy model
implementations sometimes differed by several percent at
certain wavelengths [301, 302]. These surprisingly large
discrepancies for a fundamental quantity such as the ETS
can be explained by the difficulty in measuring this spectrum. In one method, several solar spectra are observed at
the Earth’s surface at different path lengths of the direct
solar beam through the atmosphere. These measurements
are then extrapolated using the Langley technique [303] to
a path length of zero for deriving the ETS. The method is
subject to large uncertainties at wavelengths where atmospheric attenuation is large, such as at wavelengths shorter
than 310 nm (where ozone absorbs strongly) or in strong
water vapor absorption bands. Another method is the direct
measurement from space. The challenge of this method is to
prevent changes in an instrument’s calibration during transport from the calibration laboratory to space. Both methods
have advanced greatly during the last years.
Gröbner et al. [304] applied the Langley technique to
radiometrically accurate measurements of QASUME (Sec
10.1) and a “Fourier-transform spectroradiometer,” which
measures spectra at high resolution, to derive an ETS over
the wavelength range of 300–500 nm with a spectral resolution of 0.025 nm, a wavelength accuracy of 0.01 nm, and
a radiometric accuracy of 2% (95% CL) between 310 and
500 nm and 4% at 300 nm. Richard et al. [305] measured
the ETS from the International Space Station with the Total
and Spectral Solar Irradiance Sensor / Spectral Irradiance
Monitor (TSIS-1 SIM) between 200 and 2,400 nm with an
accuracy of 0.5% (95% CL) and a spectral resolution of 5 nm
between 280 and 400 nm. The high accuracy is achieved
by calibrating the system against a cryogenic radiometer
and monitoring the instrument’s stability in space with an
on-board, detector-based reference electrical substitution
radiometer. Finally, by combining the superior spectral
13
970
resolution of the spectrum by Gröbner et al. [304] with the
greater radiometric accuracy of the TSIS-1 SIM spectrum,
Coddington et al. [306] developed a composite spectrum
(named TSIS-1 HSRS) with a spectral resolution of 0.025,
a sampling resolution of 0.01 nm and a radiometric accuracy of better than 1.3% (68% CL) at wavelengths shorter
than 400 nm, representative of solar minimum conditions
between solar cycles 24 and 25. This spectrum can be considered a new benchmark for modeling applications.
An important application of radiative transfer models is
the calculation of UV irradiances at the Earth’s surface from
backscattered radiances measured by satellites (Sect. 10.3).
Typically, measurements at different wavelengths by a single
space-based instrument such as OMI are used to first derive
the TCO and then apply corrections to account for the effects
of clouds and aerosols [307].
10.3 Satellite observations of UV radiation
The TCO and UV radiation at the ground have been estimated from measurements of various space-borne sensors
since the 1970s, starting with the Backscatter Ultraviolet
(BUV) experiment on the Nimbus-4 satellite [308]. These
measurements have been continued, amongst others, by
several Solar Backscatter UV (SBUV) instruments [309];
Total Ozone Monitoring Spectrometers (TOMS) [310,
311]; Global Ozone Monitoring Experiments (GOME and
GOME-2) [312, 313]; the Ozone Monitoring Instrument
(OMI) [314] on the Aura satellite; and the Earth Polychromatic Imaging Camera (EPIC) installed on the Deep Space
Climate Observatory (DSCOVR), which is located at the
Lagrange Point L1 between the Earth and Sun [315].
Several of these types of instruments have been installed
on various satellites. Estimates of UV radiation are derived
from backscattered radiances measured by these sensors and
radiative transfer model calculations (Sect. 10.2). Uncertainties of these estimates are typically larger than those of
UV measurements at the Earth’s surface because the conditions on the ground cannot be completely characterized
from space, in particular in the presence of clouds [316],
absorbing aerosols in the boundary layer [317], or snow and
ice [318]. The validation of satellite data with ground-based
measurements from many sites has been discussed in our
previous assessment [9]. In general, UV data from satellites are accurate within a few percent under low-aerosol
and clear-sky conditions, but can be affected by systematic
errors exceeding 50% for less ideal observing conditions.
Data of UV radiation at the Earth’s surface estimated
from satellite observations typically have the spatial resolution of the satellite sensor (e.g., 13 × 24 km at nadir for
OMI) and are typically based on one satellite-measured
spectrum per day at low and mid-latitudes. As an alternative, Kosmopoulos et al. [319] have used inputs from various
13
Photochemical & Photobiological Sciences (2023) 22:937–989
data sources to calculate real time and forecasted UVIs for
Europe with a spatial and temporal resolution of 5 km and
15 min, respectively. The new data product agrees with
measurements at 17 ground-based stations distributed across
Europe to within ± 0.5 UVI units for 80% of clear-sky and
70% of all-sky conditions. Similarly, Vuilleumier et al. [320]
calculated erythemal irradiance for Switzerland with a spatial resolution of 1.5–2 km and a temporal resolution of one
hour for 2004–2018, using data from several European satellites. A validation of these data with ground-based measurements at three meteorological stations in Switzerland
(Locarno, Payerne, and Davos) indicates that the expanded
uncertainty of hourly UVI values of the new data products
is about 0.3 UVI units for UVI < 3 and up to 1.5 UVI units
for UVI > 6.
Measurements with OMI started in 2004 and their quality
has degraded recently [247]. The future of the Aura spacecraft is uncertain beyond 2023 [321]. Fortunately, several
alternative satellite instruments have become operational
within the last years to continue monitoring of ozone and
UV radiation from space. For example, the Ozone Mapping
and Profiler Suite (OMPS) [322] is installed on NOAA’s
Suomi NPP (launched in 2011) and the NOAA-20 (launched
in 2017) satellites. The TROPOspheric Monitoring Instrument (TROPOMI) [323], which is installed on the Sentinel-5 Precursor satellite (launched in 2017), will continue
ozone-monitoring efforts by the European Space Agency.
TROPOMI may also fly on future Sentinel satellites [324].
TROPOMI observations of UV radiation have recently been
compared with ground-based measurements at 25 sites
[325]. For snow-free surface conditions, the median relative
difference between UVI measurements by TROPOMI and
these ground stations was within ± 10% at 18 of 25 sites. For
10 sites, the agreement was at the ± 5% level. These differences are comparable to those reported for OMI [316, 318,
326, 327]. Larger differences were observed at locations
with challenging conditions, such as mountainous areas or
sites in the Arctic and Antarctic with variable snow cover.
A comprehensive comparison between OMI and TROPOMI
surface UV products is planned [314] to ensure that there is
no step-change in the time series of UV radiation measurements when transitioning from OMI to TROPOMI.
In preparation for new satellite missions (e.g., Sentinel-4
and Sentinel-5 of the European Space Agency), Lipponen
et al. [328] developed an approach to assimilate input data
from geosynchronous and low Earth orbit satellite measurements with the goal to provide high-resolution UVI and
UV-A data. Zhao and He [329] combined TCO data from
OMI with top-of-the-atmosphere reflectance data from
MODIS for quantifying attenuation by clouds and aerosols and surface reflectance data from MODIS and used a
machine learning algorithm to calculate erythemal irradiances at 1 km resolution. The system is trained and tested
Photochemical & Photobiological Sciences (2023) 22:937–989
with UV measurements of NOAA’s Surface Radiation
Budget Network (SURFRAD) and UV data from the United
States Department of Agriculture’s (USDA) UV-B Monitoring and Research Program. For most stations, calculated and
measured data agreed to within ± 5% (mean bias calculated
from match-up data). However, the system was trained with
data from the continental United States only, and the fidelity
of the method for sites that are different in terms of latitude,
ozone climatology, pollution levels, and surface albedo has
not yet been demonstrated.
10.4 Forecasting of the UV Index
The UVI is now part of weather forecasts in many countries.
National weather services and other agencies use models
to predict the diurnal course of the UVI (e.g., every hour)
for one or several days into the future (e.g., the Israel Meteorological Service (https://ims.gov.il/en/UVIHourly), the
German Meteorological Service (https:// kunden. dwd. de/
uvi/index.jsp), and the Copernicus Atmosphere Monitoring
Service (https://climate-adapt.eea.europa.eu/obser vatory/
evide nce/ proje ctions- and- tools/ cams- uv- index- forec ast).
New methods for improving UVI forecasts have recently
been proposed based on an “ensemble member” approach,
where a model is executed multiple times with different
initial conditions [330], and a machine learning algorithm
[331].
10.5 Personal exposure
Our 2014 and 2018 assessments [9, 162] discussed advances
in the understanding of personal exposure to ambient solar
UV radiation and how personal exposure relates to measurements of UV irradiance, which are typically referenced
to a horizontal surface. Exposure studies address needs for
both research and public advice and quantify UV radiation
on non-horizontal surfaces, and how the effects of shade,
clothing, and human behavior affect UV doses in real-world
settings. Exposure studies have shown that adults working
outdoors receive only about 10% of the total available annual
UV radiation dose, while indoor-working adults and children
get only about 2–4% of the available UV dose [332, 333].
This shows that standard irradiance measurements are a poor
proxy for realistic exposures. While there could be a good
correlation between ambient and personal UV dose at the
population level, exposure of individuals depends greatly on
lifestyle. Reviews of a large number of studies on personal
exposure to UV radiation during non-occupational [334] and
occupational [335] activities concluded that understanding
of human exposure to UV radiation has greatly increased
during the last 4–5 decades. However, for most activities, our ability to accurately calculate the UV exposure of
exposed body sites is still limited for many conditions.
971
10.5.1 Exposure models
Models of human morphology can quantify the protection afforded by attire, for example, from wearing various
hats [336] and sunglasses [337]. These models often use
the “predictive protection factor” (PPF), which is akin to
the sun protection factor (SPF) developed for sunscreens,
except that the PPF also depends on the direct-to-diffuse
ratio of incident radiation. These models may be validated
using mannequin torsos or heads equipped with UV sensors
[338]. The sky view factor derived from all-sky imagery in
the visible range together with the calculated clear-sky UV
irradiance has recently been utilized to accurately estimate
UV irradiance in partially shaded settings [339].
Doses of erythemal radiation received by the human body
during holidays at the beach have recently been modeled
[340]. Taking into account all confounding factors affecting
exposure (e.g., clothing, behavior, photo-protection), these
models predict that the forearm typically receives about 170
standard erythemal doses (SED) in a week, which is comparable with the average annual exposure of a citizen in Europe
or North America. Furthermore, for a full day sun-bathing
at the beach or pool, multiple body sites can receive more
than 50 SED.
10.5.2 Personal dosimetry
The three types of dosimeters previously identified [162]—
polysulphone (a plastic film that changes its transmission
following exposure to UV radiation), biofilm, and electronic devices—are still in use, and their relative merits in
different contexts have recently been reviewed [341, 342].
These measurement technologies were further described in
a review that also proposes a future course for development
and regulation of wearable UV sensors [343].
Some authors [e.g., 344] distinguish between “radiometers,” which give an instantaneous flux reading such as the
UVI, and “dosimeters,” which measure cumulative dose
such as the standard erythemal dose (SED). However, the
distinction is irrelevant for many electronic sensors, which
measure flux but also accumulate it electronically. The same
can apply to photochromic sensors in combination with
smartphones or other electronic logging. Hereafter, we use
the term “dosimeter” for all types of sensors.
The history and characteristics of polysulphone dosimeters have been reviewed by one of their pioneers [285]. They
are useful whenever water resistance is necessary, as in a
study of triathletes [286]. Alternative photochromic sensors
13
972
have been developed using the photodegradable dye DTEC37
[345] and xanthomattin [344].
A new development of a biofilm dosimeter that mimics the photoreaction resulting in previtamin D3 synthesis
in human skin has recently been presented [346]. Biofilm
sensors of a similar type were used to measure exposure to
UV radiation of lifeguards, demonstrating that this group
receives high doses of erythemal UV radiation, averaging
over 6 SEDs per day [347].
Electronic dosimeters have some advantages for research
involving personal dosimetry compared to other sensors.
They can be engineered to have a spectral responsivity
and a directional response approaching those of researchgrade radiometers measuring erythemal irradiance [284].
The time resolution and ability to interface wirelessly with
smartphones allows feedback to users, and has supported
research on how such information can influence sun exposure amongst melanoma survivors [348], dockworkers and
fishermen [349], or young adults in general [350]. In a small
study of outdoor workers in Romania, dosimeters measured
up to 6 SEDs per day and led the authors to suggest that UV
dosimeters should be compulsory for outdoor workers, similar to personal dosimetry for ionizing radiation in relevant
professions [351].
A 14-year study with electronic dosimeters showed that
participants that are in continued employment maintained
their sun exposure behavior, retirees increased their exposure, and high school students reduced their exposure when
starting work [352]. Additional exposure studies confirmed
expectations that outdoor workers [351]; participants in triathlons [286]; and elite surfers, windsurfers, and Olympic
sailors [353] are at high risk of overexposure to UV radiation. In general, staying outdoors for long periods, even at
low UV irradiance levels, can result in risk of damage from
UV radiation [232].
Airline pilots have long been known to have twice the
incidence rate of malignant melanoma and keratinocyte skin
cancers than the general population, but UV-B radiation is
almost entirely blocked by cockpit windows [354]. Other
factors explaining this elevated risk of skin cancer, like ionizing radiation and disrupted circadian rhythms, have been
largely ruled out. Measurements with dosimeters that are
sensitive to both erythemal and UV-A radiation suggested
that cockpit windows are partially transparent to UV-A
radiation and pilots are therefore exposed to levels of UV-A
radiation that exceed guidelines for eye protection established by ICNIRP [355], in particular if sunglasses are not
worn or visors are not deployed [356].
37
(2Z,6Z)-2,6-bis(2-(2,6-diphenyl-4H-thiopyran-4-ylidene)ethylidene)cyclohexanone.
13
Photochemical & Photobiological Sciences (2023) 22:937–989
10.5.3 Low‑cost/crowd‑sourced sensors and cell phone
apps
Our last assessment [9] described a wide range of new tools
for research and for getting information to users, including electronic sensors, photochromic films with associated
software, and forecasts or “nowcasts” of UV radiation using
cell phone apps. A review of developments in this area [357]
describes the promise of these new technologies, but a comparison of UV radiation reported by cell phone apps with
actual UV measurements found that many of these apps have
poor accuracy [358]. For example, of the six apps reviewed
in this study, only one was able to predict the actual UVI
to within ± 30% in most cases. A further miniaturization of
sensors to millimeter scale with wireless communication to
standard consumer devices [359] will widen the scope of
how these sensors can be deployed. Other studies have also
shown that useful personal exposures to UV radiation can be
achieved from satellite-based UV radiation estimates combined with exposure ratio modeling to account for individual
factors [360] or by leveraging UV data from local research
stations [361].
11 Action spectra
Action spectra describe the wavelength dependence of biological effects caused by UV radiation. A biological effect
is quantified by first multiplying the action spectrum for this
effect by the spectrum of the incident irradiance and then
integrating this product over wavelength. The result is the
biologically effective UV irradiance, UVBE. Most action
spectra decrease by several orders of magnitude towards
longer wavelengths in the UV-B range. Since solar spectra
increase by a similar amount in this wavelength range, a
given biological effect is very sensitive to the wavelength
intervals within the UV-B range over which this decrease
(action spectrum) or increase (solar spectrum) occurs.
This implies that action spectra must be very accurately
measured.
The most widely used action spectrum is that for erythema [10], which is the basis of UVI calculations. In sunlight, the strongest contribution to erythema is from UV-B
wavelengths, peaking near 307 nm. UV-A wavelengths also
contribute, especially at the shorter end of the UV-A region
(e.g., 315–340 nm). A small-scale study with 10 participants
[362] found clinically perceptible erythema after exposure
to UV radiation in the 370–400 nm range plus visible light
(400–700 nm), confirming that longer UV-A wavelengths
can also cause erythema. The study also suggests that the
erythema action spectrum, which is currently defined only
up to 400 nm [10], should possibly be extended into the
Photochemical & Photobiological Sciences (2023) 22:937–989
973
visible range. This finding is also supported by a recent
assessment by Diffey and Osterwalder [363].
Another important action spectrum for human health
defines the wavelength dependence of the conversion of
7-dehydrocholesterol in the skin to previtamin D3, which
is subsequently transformed to the active form of vitamin D
(1,25-dihydroxycholecalciferol or calcitriol) involving isomerisation and hydroxylations in the skin, liver, and kidneys.
This spectrum was measured 40 years ago [364] and was
standardized by the International Commission on Illumination (CIE) [365] by interpolating the original data, plus
extending the end of the spectrum from 315 to 330 nm via
an exponential extrapolation. The spectrum has been widely
used for developing recommendations for optimal solar
exposure [179]; however, its validity has been questioned
[179, 366]. Specifically, the CIE standard [365] is based
on a scanned figure from a single publication that does not
include a complete description of the experiment such as the
UV doses used. Furthermore, the source used for irradiation
had a large bandwidth of 5 nm, which leads to noticeable
broadening of the spectrum, and the extrapolation from 315
to 330 nm is questionable because there are no experimental
data in this wavelength range.
Young et al. [367] have recently provided evidence that
shifting the CIE action spectrum for previtamin D3 synthesis
by 5 nm to shorter wavelengths (Fig. 10) would produce a
more realistic action spectrum for the production of previtamin D3 in human skin. They exposed 75 volunteers to
five lamp spectra with different spectral composition, and
correlated the observed increase in serum 25(OH)D levels (the form of vitamin D used to assess vitamin D status)
with the effective UV irradiance, UVBE. The action spectrum for calculating UVBE was either the CIE spectrum in
its unaltered form or a variant shifted in wavelength. The
shift by 5 nm is plausible because the absorption spectrum
of 7-dehydrocholesterol is also found to be shifted by about
5 nm to shorter wavelengths relative to the CIE action spectrum, even after adjusting for the spectral transmission of
0
Relative effect
10
10
10
CIE erythema action spectrum (1998)
CIE previtamin D action spectrum (2006)
CIE previtamin D action spectrum (2006),
shifted by 5 nm to shorter wavelengths
-1
-2
-3
10
10
-4
280
300
320
340
360
380
400
Wavelength [nm]
Fig. 10 Comparison of CIE action spectra for erythema [10] and the
cutaneous synthesis of previtamin D3 [365]. The effect of a 5-nm blue
shift on the previtamin D3 action spectrum is also shown
the skin’s outermost layer, the stratum corneum [366]. Furthermore, results obtained with the shifted action spectrum
are consistent with calculations using alternative vitamin D
action spectra proposed by Bolsée et al. [368], Olds [369],
and van Dijk et al. [370], which are also shifted to shorter
wavelengths relative to the CIE spectrum. These results suggest that the CIE standard [365] may need revision. However, the spectral change of solar spectra observed on the
Earth (e.g., the difference between summer at the equator
and winter in the Northern Hemisphere) is smaller than the
difference in the spectral composition of the various artificial
light sources used in the new experiment. The effect of the
shift is, therefore, less important for natural sunlight, leading
to the conclusion by Young et al. [367] that the CIE action
spectrum (with no shift) remains adequate for risk–benefit
calculations and the development of recommendations for
healthy solar exposure. Along the same line, a recent assessment [371] concluded that the current CIE action spectrum
[365] probably needs to be amended, but that it is acceptable to continue using this action spectrum for risk-benefits
assessments until that work is completed.
An action spectrum for the inhibition of SARS-CoV-2
(the virus responsible for the COVID-19 disease) was
recently measured. This spectrum is discussed by Bernhard
et al. [2].
12 Gaps in knowledge
Our assessment identified the following gaps in knowledge:
• Most ODSs are also GHGs and have a large effect on
global warming. However, since ozone is also a GHG,
depletion of ozone caused by ODSs has a cooling effect
(Sect. 4.2). The net effect on temperatures at the Earth’s
surface resulting from the direct (warming) effect of
ODSs and the indirect (cooling) effect from ozone depletion induced by ODSs is uncertain because climate models disagree on the magnitude of the latter effect. While
the balance of all studies suggests that the Montreal Protocol is highly effective in limiting temperature rise at
the Earth’s surface, the magnitude of the effect remains
uncertain.
• The effect of Antarctic ozone depletion on changes in sea
ice surrounding Antarctica is not well understood.
• The effect of the Antarctic ozone hole on summertime
weather in the Southern Hemisphere is uncertain. In particular, it is difficult to quantify if changes in weather
are more affected by the year-to-year variability of the
polar vortex, which is partly driven by changes in sea
surface temperature of the Southern Ocean, or by the
actual depletion of ozone within the vortex. It is also
not clear how the coupling between the stratosphere and
13
974
•
•
•
•
•
•
Photochemical & Photobiological Sciences (2023) 22:937–989
troposphere in weak vortex conditions will evolve under
ozone recovery.
While several studies have identified correlations
between Arctic ozone changes and weather in the Northern Hemisphere, knowledge on how these linkages are
mediated is incomplete.
The paucity of measurements of the properties of aerosols in the UV-B range hampers our ability to accurately
assess the effects of aerosols on a global scale as well as
for urban regions. While efforts to improve this situation
are underway—for example, EUBREWNET has recently
started to provide AOD in the wavelength range from 306
to 320 nm (Sect. 6.1)—aerosol data in the UV-B range
are currently available only for a few locations.
Atmospheric blocking systems (stagnant high- or lowpressure synoptic systems) can cause week-long anomalies of UV radiation. It is not well understood how climate change may alter the frequency, persistence, and
geographical extent and location of these blocking patterns, and their effect on UV radiation.
One of the largest uncertainties in projecting changes to
ozone and UV radiation during the twenty-first century is
the evolution of GHG trajectories, which mostly depend
on policy decisions and societal behavior.
Uncertainties in projections of UV radiation arising from
incomplete knowledge of future changes in aerosol and
cloud optical properties are significant.
The number of stations with high-quality spectral UV
measurements has been declining during the last decade and the funding for many of the remaining stations
is uncertain. If this trend continues, the scientific community may lose the ability to assess changes of UV
radiation at the Earth’s surface and associated impacts,
in order to verify new satellite UV data products with
ground-based observations and to validate model projections.
13 Conclusions
Virtually all studies published during the last four years
confirmed that changes in UV radiation (typically assessed
with the UVI) during the last 25 years have been small: less
than 4% per decade for the UVI at the majority of ground
stations, increasing at some sites and decreasing at others.
Changes in the UVI outside the polar regions over the last
2–3 decades were mainly governed by variations in clouds,
aerosols, and surface reflectivity (for snow- or ice-covered
areas), while changes in TCO are less important. Variability in the UVI in Antarctica continued to be very large. In
spring 2019, the UVI was at the minimum of the historical
(1991–2018) range at the South Pole, while near record-high
values were observed in spring 2020 and 2021, which were
13
up to 80% above the historical mean. In the Arctic, some
of the highest UV-B irradiances on record were measured
in March and April 2020. For example in March 2020, the
monthly average UVI over the Canadian Arctic was up to
70% higher than the historical (2005–2019) average, often
exceeding this mean by three standard deviations.
Without the Montreal Protocol, the UVI at northern and
southern latitudes of less than 50° would have increased
by 10–20% between 1996 and 2020. For southern latitudes
exceeding 50°, the UVI would have surged by between 25%
(year-round at the southern tip of South America) and more
than 100% (South Pole in spring).
Under the presumption that all countries will adhere to
the Montreal Protocol in the future and that atmospheric
aerosol concentrations remain constant, the UVI at midlatitudes (30–60°) is projected to decrease between 2015
and 2090 by 2–5% in the north and by 4–6% in the south
due to recovering ozone. Changes projected for the tropics
are smaller than 3%.
Since most substances controlled by the Montreal Protocol are also greenhouse gases, the phase-out of these substances may have avoided warming by 0.5 to 1.0 °C over
mid-latitude regions of the continents, and by more than
1.0 °C in the Arctic. ODSs contributed one-half of the forced
Arctic sea ice loss in the latter half of the twentieth century.
The uncertainty of changes in temperature and sea ice simulated by these models is still large.
Assessing the Montreal Protocol’s impact on solar UV
radiation and climate, and their interaction, is impeded by
several gaps in knowledge. The net temperature change at the
Earth’s surface resulting from the direct (warming) effect of
ODSs and the indirect (cooling) effect from ozone depletion
is uncertain, because climate models disagree on the magnitude of the latter effect. While all studies support the role of
the Montreal Protocol in limiting global warming, the magnitude of increases in temperatures that were averted remains
uncertain. There is evidence that in both hemispheres polar
ozone depletion in spring has an influence on weather; however, the mechanisms and magnitude of the effect are not fully
understood. The lack of measurements of absorption properties of aerosols in the UV-B range hinders the assessment of
the aerosols’ impact on UV-B radiation. One of the largest
uncertainties in projecting changes in UV radiation during
the twenty-first century is the incomplete knowledge of how
GHGs will increase over time. Uncertainties in UV projections arising from inadequate understanding of future changes
in aerosols and clouds are also significant.
Our assessment addresses several United Nations Sustainable Development Goals (SDGs) and their targets (https://sdgs.
un.org/goals). Owing to the Montreal Protocol, large increases
in UV-B radiation have been avoided and global warming
reduced. By assessing how ozone depletion affects climate
change, we contribute to SDGs 13.1 (“strengthen resilience
Photochemical & Photobiological Sciences (2023) 22:937–989
to climate-related hazards and disasters”) and 13.2 (“integrate
climate change measures into policy, strategy and planning”).
Furthermore, by providing up-to-date information on the interactive effects of ozone depletion on UV radiation and climate,
both in this assessment and the companion document titled
“Questions and Answers about the effects of the depletion of
the ozone layer on humans and the environment”, we address
SDGs 13.3 (“improve education on climate-change mitigation”) and 17.14 (“enhance policy coherence for sustainable
development”).
Acknowledgements Generous contributions by UNEP/Ozone Secretariat were provided for the convened author meeting. GHB acknowledges travel funding provided by the U.S. Global Change Research
Program. AFB’s contribution was partly supported by research funds of
the Laboratory of Atmospheric Physics, Aristotle University of Thessaloniki, Greece. Figures 1, 2, 3, 6, 8, and 9 were reprinted or adapted
from sources published under the Creative Commons Attribution 4.0
International License (CC BY 4.0; https://creativecommons.org/licen
ses/by/4.0/).
Author contributions All authors contributed to the conception and
assessment and carried out extensive revisions of content.
Funding Open access funding provided by HEAL-Link Greece. Open
access funding was provided by Aristotle University, Thessaloniki,
Greece.
Data availability All data generated or analyzed are either included in
this published article or part of the analyses of papers cited.
Declarations
Conflict of interest The authors have no conflicts of interest.
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/.
References
1. Neale, R. E., Lucas, R. M., Byrne, S., Hollestein, L., Rhodes, L.
E., Yasar, S., Young, A. R., Berwick, M., Ireland, R., & Olsen,
C. M. (2023). The effects of exposure to solar radiation on human
health. Photochemical & Photobiological Sciences. https://doi.
org/10.1007/s43630-023-00375-8
2. Bernhard, G. H., Madronich, S., Lucas, R. M., Byrne, S. N.,
Schikowski, T., & Neale, R. E. (2023). Linkages between
COVID-19, solar UV radiation, and the Montreal Protocol. Photochemical & Photobiological Sciences. https://doi.org/10.1007/
s43630-023-00373-w
975
3. Barnes, P. W., Robson, T. M., Zepp, R. G., Bornman, J. F.,
Jansen, M. A. K., Ossola, R., Wang, Q.-W., Robinson, S. A., Foereid, B., Klekociuk, A. R., Martinez-Abaigar, J., Hou, W. C., &
Paul, N. D. (2023). Interactive effects of changes in UV radiation
and climate on terrestrial ecosystems, biogeochemical cycles,
and feedbacks to the climate system. Photochemical & Photobiological Sciences. https://doi.org/10.1007/s43630-023-00376-7
4. Neale, P. J., Williamson, C. E., Banaszak, A. T., Häder, D. P.,
Hylander, S., Ossola, R., Rose, K. C., Wängberg, S.-Å., & Zepp,
R. G. (2023). The response of aquatic ecosystems to the interactive effects of stratospheric ozone depletion, UV radiation, and
climate change. Photochemical & Photobiological Sciences.
https://doi.org/10.1007/s43630-023-00370-z
5. Madronich, S., Sulzberger, B., Longstreth, J. D., Schikowski,
T., Sulbæk Andersen, M. P., Solomon, K. R., & Wilson, S. R.
(2023). Changes in tropospheric air quality related to the protection of stratospheric ozone in a changing climate. Photochemical & Photobiological Sciences. https://doi.org/10.1007/
s43630-023-00369-6
6. Andrady, A. L., Heikkila, A. M., Pandey, K. K., Bruckman, L. S.,
White, C. C., Zhu, M., & Zhu, L. (2023). Effects of UV radiation
on natural and synthetic materials. Photochemical & Photobiological Sciences. https://doi.org/10.1007/s43630-023-00377-6
7. Jansen, M. A. K., Barnes, P. W., Bornman, J. F., Rose, K. C.,
Madronich, S., White, C. C., Zepp, R. G., & Andrady, A. L.
(2023). The Montreal Protocol and the fate of environmental
plastic debris. Photochemical & Photobiological Sciences.
https://doi.org/10.1007/s43630-023-00372-x
8. United Nations. (1987). Montreal Protocol on Substances that
Deplete the Ozone Layer. United Nations Treaty Series, Number 1522. https://treaties.un.org/Pages/showDetails.aspx?objid=
080000028003f7f7
9. Bais, A. F., Bernhard, G., McKenzie, R. L., Aucamp, P. J.,
Young, P. J., Ilyas, M., Jockel, P., & Deushi, M. (2019). Ozoneclimate interactions and effects on solar ultraviolet radiation.
Photochemical & Photobiological Sciences, 18(3), 602–640.
https://doi.org/10.1039/c8pp90059k
10. CIE. (1998). Erythema reference action spectrum and standard
erythema dose. CIE Standard Bureau, Vol. ISO 17166:1999(E),
CIE DS 007.1/E-1998. Commission Internationale de
l’Eclairage, Vienna, Austria
11. WMO. (2022). Scientific Assessment of Ozone Depletion: 2022,
GAW Report No. 278. World Meteorological Organization,
Geneva, Switzerland.
12. Weber, M., Arosio, C., Coldewey-Egbers, M., Fioletov, V. E.,
Frith, S. M., Wild, J. D., Tourpali, K., Burrows, J. P., & Loyola, D. (2022). Global total ozone recovery trends attributed to
ozone-depleting substance (ODS) changes derived from five
merged ozone datasets. Atmospheric Chemistry and Physics,
22(10), 6843–6859. https://doi.org/10.5194/acp-22-6843-2022
13. WMO. (2018). Scientific Assessment of Ozone Depletion: 2018,
Global Ozone Research and Monitoring Project – Report No. 58.
World Meteorological Organization, Geneva, Switzerland.
14. Steinbrecht, W., Hegglin, M. I., Harris, N., & Weber, M. (2018).
Is global ozone recovering? Comptes Rendus Geoscience, 350(7),
368–375. https://doi.org/10.1016/j.crte.2018.07.012
15. Solomon, S., Ivy, D. J., Kinnison, D., Mills, M. J., Neely, R. R.,
& Schmidt, A. (2016). Emergence of healing in the Antarctic
ozone layer. Science, 353(6296), 269–274. https://doi.org/10.
1126/science.aae0061
16. Kuttippurath, J., Kumar, P., Nair, P. J., & Pandey, P. C. (2018).
Emergence of ozone recovery evidenced by reduction in the
occurrence of Antarctic ozone loss saturation. npj Climate
and Atmospheric Science, 1(1), 42. https:// doi. org/ 10. 1038/
s41612-018-0052-6
13
976
17. Pazmiño, A., Godin-Beekmann, S., Hauchecorne, A., Claud,
C., Khaykin, S., Goutail, F., Wolfram, E., Salvador, J., & Quel,
E. (2018). Multiple symptoms of total ozone recovery inside
the Antarctic vortex during austral spring. Atmospheric Chemistry and Physics, 18(10), 7557–7572. https://doi.org/10.5194/
acp-18-7557-2018
18. Tully, M. B., Krummel, P. B., & Klekociuk, A. R. (2019). Trends
in Antarctic ozone hole metrics 2001–17. Journal of Southern
Hemisphere Earth Systems Science, 69(1), 52–56. https://doi.org/
10.1071/es19020
19. Bodeker, G. E., & Kremser, S. (2021). Indicators of Antarctic
ozone depletion: 1979 to 2019. Atmospheric Chemistry and Physics, 21(7), 5289–5300. https://doi.org/10.5194/acp-21-5289-2021
20. Kramarova, N., Newman, P. A., Nash, E. R., Strahan, S. E., Long,
C. S., Johnson, B., Pitts, M., Santee, M. L., Petropavlovskikh, I.,
Coy, L., de Laat, J., Bernhard, G. H., Stierle, S., & Lakkala, K.
(2021). Antarctic ozone hole. In State of the Climate in 2020,
Bulletin of the American Meteorological Society. https://doi.org/
10.1175/BAMS-D-21-0081.1
21. Zambri, B., Solomon, S., Thompson, D. W. J., & Fu, Q. (2021).
Emergence of Southern Hemisphere stratospheric circulation
changes in response to ozone recovery. Nature Geoscience,
14(9), 638–644. https://doi.org/10.1038/s41561-021-00803-3
22. Kramarova, N., Newman, P. A., Nash, E. R., Strahan, S. E., Long,
C. S., Johnson, B., Pitts, M., Santee, M. L., Petropavlovskikh, I.,
Coy, L., & de Laat, J. (2020). Antarctic ozone hole. In State of
the Climate in 2019, Bulletin of the American Meteorological
Society. https://doi.org/10.1175/BAMS-D-20-0090.1
23. Milinevsky, G., Evtushevsky, O., Klekociuk, A., Wang, Y.,
Grytsai, A., Shulga, V., & Ivaniha, O. (2019). Early indications
of anomalous behaviour in the 2019 spring ozone hole over
Antarctica. International Journal of Remote Sensing, 41(19),
7530–7540. https://doi.org/10.1080/2150704X.2020.1763497
24. Shen, X., Wang, L., & Osprey, S. (2020). Tropospheric forcing of
the 2019 Antarctic sudden stratospheric warming. Geophysical
Research Letters. https://doi.org/10.1029/2020gl089343
25. Yamazaki, Y., Matthias, V., Miyoshi, Y., Stolle, C., Siddiqui,
T., Kervalishvili, G., Laštovička, J., Kozubek, M., Ward, W.,
Themens, D. R., Kristoffersen, S., & Alken, P. (2020). September 2019 Antarctic sudden stratospheric warming: quasi-6-Day
wave burst and ionospheric effects. Geophysical Research Letters. https://doi.org/10.1029/2019gl086577
26. Safieddine, S., Bouillon, M., Paracho, A. C., Jumelet, J., Tencé,
F., Pazmino, A., Goutail, F., Wespes, C., Bekki, S., Boynard,
A., Hadji-Lazaro, J., Coheur, P. F., Hurtmans, D., & Clerbaux,
C. (2020). Antarctic ozone enhancement during the 2019 sudden stratospheric warming event. Geophysical Research Letters.
https://doi.org/10.1029/2020gl087810
27. Stone, K. A., Solomon, S., Kinnison, D. E., & Mills, M. J.
(2021). On recent large Antarctic ozone holes and ozone recovery metrics. Geophysical Research Letters. https://doi.org/10.
1029/2021gl095232
28. Yook, S., Thompson, D. W. J., & Solomon, S. (2022). Climate
impacts and potential drivers of the unprecedented Antarctic
ozone holes of 2020 and 2021. Geophysical Research Letters.
https://doi.org/10.1029/2022gl098064
29. Barnes, P. W., Robson, T. M., Neale, P. J., Williamson, C. E.,
Zepp, R. G., Madronich, S., Wilson, S. R., Andrady, A. L., Heikkilä, A. M., Bernhard, G. H., Bais, A. F., Neale, R. E., Bornman,
J. F., Jansen, M. A. K., Klekociuk, A. R., Martinez-Abaigar, J.,
Robinson, S. A., Wang, Q. W., Banaszak, A. T., … Young, A.
R. (2022). Environmental effects of stratospheric ozone depletion, UV radiation, and interactions with climate change: UNEP
Environmental Effects Assessment Panel, Update 2021. Photochemical & Photobiological Sciences.https://doi.org/10.1007/
s43630-022-00176-5
13
Photochemical & Photobiological Sciences (2023) 22:937–989
30. Zhuo, Z., Kirchner, I., Pfahl, S., & Cubasch, U. (2021). Climate
impact of volcanic eruptions: The sensitivity to eruption season
and latitude in MPI-ESM ensemble experiments. Atmospheric
Chemistry and Physics, 21(17), 13425–13442. https://doi.org/
10.5194/acp-21-13425-2021
31. Klekociuk, A. R., Tully, M. B., Krummel, P. B., Henderson, S. I.,
Smale, D., Querel, R., Nichol, S., Alexander, S. P., Fraser, P. J.,
& Nedoluha, G. (2022). The Antarctic ozone hole during 2020.
Journal of Southern Hemisphere Earth Systems Science, 72(1),
19–37. https://doi.org/10.1071/es21015
32. Ivanciu, I., Matthes, K., Wahl, S., Harlaß, J., & Biastoch, A.
(2021). Effects of prescribed CMIP6 ozone on simulating the
Southern Hemisphere atmospheric circulation response to ozone
depletion. Atmospheric Chemistry and Physics, 21(8), 5777–
5806. https://doi.org/10.5194/acp-21-5777-2021
33. Strahan, S. E., Douglass, A. R., & Damon, M. R. (2019). Why do
Antarctic ozone recovery trends vary? Journal of Geophysical
Research: Atmospheres, 124(15), 8837–8850. https://doi.org/10.
1029/2019jd030996
34. Cordero, R. R., Feron, S., Damiani, A., Redondas, A., Carrasco,
J., Sepúlveda, E., Jorquera, J., Fernandoy, F., Llanillo, P., Rowe,
P. M., & Seckmeyer, G. (2022). Persistent extreme ultraviolet
irradiance in Antarctica despite the ozone recovery onset. Scientific Reports. https://doi.org/10.1038/s41598-022-05449-8
35. Manney, G. L., Livesey, N. J., Santee, M. L., Froidevaux, L.,
Lambert, A., Lawrence, Z. D., Millán, L. F., Neu, J. L., Read,
W. G., Schwartz, M. J., & Fuller, R. A. (2020). Record-low Arctic stratospheric ozone in 2020: MLS observations of chemical
processes and comparisons with previous extreme winters. Geophysical Research Letters. https://doi.org/10.1029/2020gl089063
36. Lawrence, Z. D., Perlwitz, J., Butler, A. H., Manney, G. L., Newman, P. A., Lee, S. H., & Nash, E. R. (2020). The remarkably
strong Arctic stratospheric polar vortex of winter 2020: Links
to record-breaking Arctic oscillation and ozone loss. Journal of
Geophysical Research. https://doi.org/10.1029/2020JD033271
37. Manney, G. L., Santee, M. L., Rex, M., Livesey, N. J., Pitts,
M. C., Veefkind, P., Nash, E. R., Wohltmann, I., Lehmann, R.,
Froidevaux, L., Poole, L. R., Schoeberl, M. R., Haffner, D. P.,
Davies, J., Dorokhov, V., Gernandt, H., Johnson, B., Kivi, R.,
Kyrö, E., Larsen, N., et al. (2011). Unprecedented Arctic ozone
loss in 2011. Nature, 478(7370), 469–475. https://doi.org/10.
1038/nature10556
38. Varotsos, C. A., Efstathiou, M. N., & Christodoulakis, J. (2020).
The lesson learned from the unprecedented ozone hole in the
Arctic in 2020; a novel nowcasting tool for such extreme events.
Journal of atmospheric and solar-terrestrial physics, 207,
105330. https://doi.org/10.1016/j.jastp.2020.105330
39. Dameris, M., Loyola, D. G., Nützel, M., Coldewey-Egbers, M.,
Lerot, C., Romahn, F., & van Roozendael, M. (2021). Record low
ozone values over the Arctic in boreal spring 2020. Atmospheric
Chemistry and Physics, 21(2), 617–633. https://doi.org/10.5194/
acp-21-617-2021
40. Grooß, J. U., & Müller, R. (2021). Simulation of record arctic
stratospheric ozone depletion in 2020. Journal of Geophysical
Research. https://doi.org/10.1029/2020JD033339
41. Wohltmann, I., Gathen, P., Lehmann, R., Maturilli, M., Deckelmann, H., Manney, G. L., Davies, J., Tarasick, D., Jepsen, N.,
Kivi, R., Lyall, N., & Rex, M. (2020). Near complete local reduction of Arctic stratospheric ozone by severe chemical loss in
spring 2020. Geophysical Research Letters. https://doi.org/10.
1029/2020gl089547
42. Lee, S. H. (2021). The January 2021 sudden stratospheric warming. Weather, 76(4), 135–136. https://doi.org/10.1002/wea.3966
43. Lu, Q., Rao, J., Liang, Z., Guo, D., Luo, J., Liu, S., Wang, C., &
Wang, T. (2021). The sudden stratospheric warming in January
Photochemical & Photobiological Sciences (2023) 22:937–989
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
2021. Environmental Research Letters, 16(8), 084029. https://
doi.org/10.1088/1748-9326/ac12f4
Lee, S. H., & Butler, A. H. (2019). The 2018–2019 Arctic stratospheric polar vortex. Weather, 75(2), 52–57. https://doi.org/10.
1002/wea.3643
Butchart, N., & Remsberg, E. E. (1986). The area of the stratospheric polar vortex as a diagnostic for tracer transport on an
isentropic surface. Journal of the Atmospheric Sciences, 43(13),
1319–1339. https://doi.org/10.1175/1520-0469(1986)043<1319:
TAOTSP>2.0.CO;2
Bernhard, G., Fioletov, V., Grooß, J.-U., Ialongo, I., Johnsen,
B., Lakkala, K., Manney, G., & Müller, R. (2020). Ozone and
UV radiation. In State of the Climate in 2019, Bulletin of the
American Meteorological Society. https:// doi. org/ 10. 1175/
BAMS-D-20-0086.1
Bernhard, G., Fioletov, V., Grooß, J.-U., Ialongo, I., Johnsen,
B., Lakkala, K., Manney, G., & Müller, R. (2022). Ozone and
UV radiation. In State of the Climate in 2021, Bulletin of the
American Meteorological Society. https:// doi. org/ 10. 1175/
BAMS-D-22-0082.1
Feng, W., Dhomse, S. S., Arosio, C., Weber, M., Burrows, J. P.,
Santee, M. L., & Chipperfield, M. P. (2021). Arctic ozone depletion in 2019/20: Roles of chemistry, dynamics and the Montreal
Protocol. Geophysical Research Letters. https://doi.org/10.1029/
2020GL091911
von der Gathen, P., Kivi, R., Wohltmann, I., Salawitch, R. J., &
Rex, M. (2021). Climate change favours large seasonal loss of
Arctic ozone. Nature Communications, 12(1), 1–17. https://doi.
org/10.1038/s41467-021-24089-6
Langematz, U. (2018). Future ozone in a changing climate.
Comptes Rendus Geoscience, 350(7), 403–409. https://doi.org/
10.1016/j.crte.2018.06.015
Oberländer, S., Langematz, U., & Meul, S. (2013). Unraveling
impact factors for future changes in the Brewer-Dobson circulation. Journal of Geophysical Research. https://doi.org/10.1002/
jgrd.50775
Klobas, J. E., Weisenstein, D. K., Salawitch, R. J., & Wilmouth,
D. M. (2020). Reformulating the bromine alpha factor and equivalent effective stratospheric chlorine (EESC): Evolution of ozone
destruction rates of bromine and chlorine in future climate scenarios. Atmospheric Chemistry and Physics, 20(15), 9459–9471.
https://doi.org/10.5194/acp-20-9459-2020
Falk, S., Sinnhuber, B.-M., Krysztofiak, G., Jöckel, P., Graf, P.,
& Lennartz, S. T. (2017). Brominated VSLS and their influence
on ozone under a changing climate. Atmospheric Chemistry
and Physics, 17(18), 11313–11329. https:// doi. org/ 10. 5194/
acp-17-11313-2017
Keeble, J., Hassler, B., Banerjee, A., Checa-Garcia, R., Chiodo, G., Davis, S., Eyring, V., Griffiths, P. T., Morgenstern, O.,
Nowack, P., Zeng, G., Zhang, J., Bodeker, G., Burrows, S., Cameron-Smith, P., Cugnet, D., Danek, C., Deushi, M., Horowitz,
L. W., Kubin, A., et al. (2021). Evaluating stratospheric ozone
and water vapour changes in CMIP6 models from 1850 to 2100.
Atmospheric Chemistry and Physics, 21(6), 5015–5061. https://
doi.org/10.5194/acp-21-5015-2021
Meinshausen, M., Nicholls, Z. R. J., Lewis, J., Gidden, M. J.,
Vogel, E., Freund, M., Beyerle, U., Gessner, C., Nauels, A.,
Bauer, N., Canadell, J. G., Daniel, J. S., John, A., Krummel, P.
B., Luderer, G., Meinshausen, N., Montzka, S. A., Rayner, P. J.,
Reimann, S., Smith, S. J., et al. (2020). The shared socio-economic pathway (SSP) greenhouse gas concentrations and their
extensions to 2500. Geoscientific Model Development, 13(8),
3571–3605. https://doi.org/10.5194/gmd-13-3571-2020
McKenzie, R., Bernhard, G., Liley, B., Disterhoft, P., Rhodes,
S., Bais, A., Morgenstern, O., Newman, P., Oman, L., Brogniez, C., & Simic, S. (2019). Success of Montreal Protocol
977
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
demonstrated by comparing high-quality UV measurements
with “World Avoided” calculations from two chemistry-climate
models. Scientific Reports, 9(1), 12332. https://doi.org/10.1038/
s41598-019-48625-z
Newman, P. A., Oman, L. D., Douglass, A. R., Fleming, E. L.,
Frith, S. M., Hurwitz, M. M., Kawa, S. R., Jackman, C. H., Krotkov, N. A., Nash, E. R., Nielsen, J. E., Pawson, S., Stolarski, R.
S., & Velders, G. J. M. (2009). What would have happened to
the ozone layer if chlorofluorocarbons (CFCs) had not been regulated? Atmospheric Chemistry and Physics, 9(6), 2113–2128.
https://doi.org/10.5194/acp-9-2113-2009
Morgenstern, O., Hegglin, M. I., Rozanov, E., O’Connor, F. M.,
Abraham, N. L., Akiyoshi, H., Archibald, A. T., Bekki, S., Butchart, N., Chipperfield, M. P., Deushi, M., Dhomse, S. S., Garcia,
R. R., Hardiman, S. C., Horowitz, L. W., Jöckel, P., Josse, B.,
Kinnison, D., Lin, M., Mancini, E., et al. (2017). Review of the
global models used within phase 1 of the Chemistry-Climate
Model Initiative (CCMI). Geoscientific Model Development,
10(2), 639–671. https://doi.org/10.5194/gmd-10-639-2017
Polvani, L. M., Previdi, M., England, M. R., Chiodo, G., & Smith,
K. L. (2020). Substantial twentieth-century Arctic warming
caused by ozone-depleting substances. Nature Climate Change,
10(2), 130–133. https://doi.org/10.1038/s41558-019-0677-4
Wigley, T. M. L. (1988). Future CFC concentrations under the
Montreal Protocol and their greenhouse-effect implications.
Nature, 335(6188), 333–335. https://doi.org/10.1038/335333a0
Myhre, G., Shindell, D., Bréon, F. M., Collins, W., Fuglestvedt,
J., Huang, J., Koch, D., Lamarque, J. F., Lee, D., Mendoza, B.,
Nakajima, T., Robock, A., Stephens, G., Takemura, T., & Zhang,
H. (2013). Anthropogenic and natural radiative forcing. Climate
change 2013: The physical science basis. Contribution of Working Group I to the fifth assessment report of the Intergovernmental Panel on Climate Change (pp. 659–740). Cambridge University Press. https://doi.org/10.1017/CBO9781107415324.018
Velders, G. J. M., Andersen, S. O., Daniel, J. S., Fahey, D. W., &
McFarland, M. (2007). The importance of the Montreal Protocol
in protecting climate. Proceedings of the National Academy of
Sciences of the United States of America, 104(12), 4814–4819.
https://doi.org/10.1073/pnas.0610328104
IPCC. (2021). In V. Masson-Delmotte, P. Zhai, A. Pirani, S. L.
Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb,
M. I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J. B. R. Matthews, T. K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, & B.
Zhou (Eds.), Climate change 2021: The physical science basis.
Contribution of working group I to the sixth assessment report
of the Intergovernmental Panel on Climate Change. Cambridge
University Press. https:// doi. org/ 10. 1017/ 97810 09157 896 (in
press).
Ramaswamy, V., Schwarzkopf, M. D., & Shine, K. P. (1992).
Radiative forcing of climate from halocarbon-induced global
stratospheric ozone loss. Nature, 355(6363), 810–812. https://
doi.org/10.1038/355810a0
Shindell, D., Faluvegi, G., Nazarenko, L., Bowman, K.,
Lamarque, J.-F., Voulgarakis, A., Schmidt, G. A., Pechony, O.,
& Ruedy, R. (2013). Attribution of historical ozone forcing to
anthropogenic emissions. Nature Climate Change, 3(6), 567–
570. https://doi.org/10.1038/nclimate1835
Morgenstern, O., O’Connor, F. M., Johnson, B. T., Zeng, G.,
Mulcahy, J. P., Williams, J., Teixeira, J., Michou, M., Nabat, P.,
Horowitz, L. W., Naik, V., Sentman, L. T., Deushi, M., Bauer,
S. E., Tsigaridis, K., Shindell, D. T., & Kinnison, D. E. (2020).
Reappraisal of the climate impacts of ozone-depleting substances. Geophysical Research Letters. https://doi.org/10.1029/
2020gl088295
Morgenstern, O., Frith, S. M., Bodeker, G. E., Fioletov, V., &
A, R. J. (2021). Reevaluation of total-column ozone trends and
13
978
68.
69.
70.
71.
72.
73.
74.
75.
76.
77.
78.
79.
Photochemical & Photobiological Sciences (2023) 22:937–989
of the effective radiative forcing of ozone-depleting substances.
Geophysical Research Letters. https://doi.org/10.1029/2021G
L095376
Checa-Garcia, R., Hegglin, M. I., Kinnison, D., Plummer, D. A.,
& Shine, K. P. (2018). Historical tropospheric and stratospheric
ozone radiative forcing using the CMIP6 database. Geophysical Research Letters, 45(7), 3264–3273. https://doi.org/10.1002/
2017GL076770
Cionni, I., Eyring, V., Lamarque, J. F., Randel, W. J., Stevenson,
D. S., Wu, F., Bodeker, G. E., Shepherd, T. G., Shindell, D. T.,
& Waugh, D. W. (2011). Ozone database in support of CMIP5
simulations: Results and corresponding radiative forcing. Atmospheric Chemistry and Physics, 11(21), 11267–11292. https://doi.
org/10.5194/acp-11-11267-2011
Skeie, R. B., Myhre, G., Hodnebrog, Ø., Cameron-Smith, P.
J., Deushi, M., Hegglin, M. I., Horowitz, L. W., Kramer, R. J.,
Michou, M., Mills, M. J., Olivié, D. J. L., Connor, F. M. O.,
Paynter, D., Samset, B. H., Sellar, A., Shindell, D., Takemura, T.,
Tilmes, S., & Wu, T. (2020). Historical total ozone radiative forcing derived from CMIP6 simulations. Npj Climate and Atmospheric Science. https://doi.org/10.1038/s41612-020-00131-0
Conley, A. J., Lamarque, J. F., Vitt, F., Collins, W. D., & Kiehl,
J. (2013). PORT, a CESM tool for the diagnosis of radiative forcing. Geoscientific Model Development, 6(2), 469–476. https://
doi.org/10.5194/gmd-6-469-2013
Chiodo, G., & Polvani, L. M. (2022). New Insights on the Radiative Impacts of Ozone-Depleting Substances. Geophysical
Research Letters. https://doi.org/10.1029/2021gl096783
IPCC. (2013). In T. F. Stocker, D. Qin, G.-K. Plattner, M. Tignor,
S. K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex, & P. M.
Midgle (Eds.), Climate change 2013: The physical science basis
Contribution of Working Group I to the fifth assessment report
of the Intergovernmental Panel on Climate Change. Cambridge
University Press.
Goyal, R., England, M. H., Sen Gupta, A., & Jucker, M. (2019).
Reduction in surface climate change achieved by the 1987 Montreal Protocol. Environmental Research Letters. https://doi.org/
10.1088/1748-9326/ab4874
Liang, Y.-C., Polvani, L. M., Previdi, M., Smith, K. L., England,
M. R., & Chiodo, G. (2022). Stronger Arctic amplification from
ozone-depleting substances than from carbon dioxide. Environmental Research Letters. https:// doi. org/ 10. 1088/ 1748- 9326/
ac4a31
United Nations. (2016). Amendment to the Montreal Protocol
on Substances that Deplete the Ozone Layer. United Nations
Treaty Series, Number C.N.872.2016.TREATIES-XXVII.2.f.
https://treaties.un.org/doc/Treaties/2016/10/20161015%200323%20PM/Ch_XXVII-2.f-English%20and%20French.pdf
Velders, G. J. M., Daniel, J. S., Montzka, S. A., Vimont, I.,
Rigby, M., Krummel, P. B., Muhle, J., O’Doherty, S., Prinn, R.
G., Weiss, R. F., & Young, D. (2022). Projections of hydrofluorocarbon (HFC) emissions and the resulting global warming based
on recent trends in observed abundances and current policies.
Atmospheric Chemistry and Physics, 22(9), 6087–6101. https://
doi.org/10.5194/acp-22-6087-2022
Montzka, S. A., Dutton, G. S., Yu, P., Ray, E., Portmann, R.
W., Daniel, J. S., Kuijpers, L., Hall, B. D., Mondeel, D., Siso,
C., Nance, J. D., Rigby, M., Manning, A. J., Hu, L., Moore,
F., Miller, B. R., & Elkins, J. W. (2018). An unexpected and
persistent increase in global emissions of ozone-depleting
CFC-11. Nature, 557(7705), 413–417. https://doi.org/10.1038/
s41586-018-0106-2
Rigby, M., Park, S., Saito, T., Western, L. M., Redington, A.
L., Fang, X., Henne, S., Manning, A. J., Prinn, R. G., Dutton,
G. S., Fraser, P. J., Ganesan, A. L., Hall, B. D., Harth, C. M.,
Kim, J., Kim, K. R., Krummel, P. B., Lee, T., Li, S., Liang, Q.,
13
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
90.
91.
92.
et al. (2019). Increase in CFC-11 emissions from eastern China
based on atmospheric observations. Nature, 569(7757), 546–550.
https://doi.org/10.1038/s41586-019-1193-4
Dhomse, S. S., Feng, W., Montzka, S. A., Hossaini, R., Keeble,
J., Pyle, J. A., Daniel, J. S., & Chipperfield, M. P. (2019). Delay
in recovery of the Antarctic ozone hole from unexpected CFC-11
emissions. Nature Communications, 10(1), 5781. https://doi.org/
10.1038/s41467-019-13717-x
Lickley, M., Solomon, S., Fletcher, S., Velders, G. J. M., Daniel, J., Rigby, M., Montzka, S. A., Kuijpers, L. J. M., & Stone,
K. (2020). Quantifying contributions of chlorofluorocarbon
banks to emissions and impacts on the ozone layer and climate.
Nature Communications, 11(1), 1380. https://doi.org/10.1038/
s41467-020-15162-7
Park, S., Western, L. M., Saito, T., Redington, A. L., Henne, S.,
Fang, X., Prinn, R. G., Manning, A. J., Montzka, S. A., Fraser, P.
J., Ganesan, A. L., Harth, C. M., Kim, J., Krummel, P. B., Liang,
Q., Mühle, J., O’Doherty, S., Park, H., Park, M.-K., Reimann,
S., et al. (2021). A decline in emissions of CFC-11 and related
chemicals from eastern China. Nature, 590(7846), 433–437.
https://doi.org/10.1038/s41586-021-03277-w
Chipperfield, M. P., Hegglin, M. I., Montzka, S. A., Newman, P.
A., Park, S., Reimann, S., Rigby, M., Stohl, A., Velders, G. J. M.,
& Walter-Terrinoni, H. (2021). Report on unexpected emissions
of CFC-11. World Meteorological Organization. Switzerland.
Montzka, S. A., Dutton, G. S., Portmann, R. W., Chipperfield, M.
P., Davis, S., Feng, W., Manning, A. J., Ray, E., Rigby, M., Hall,
B. D., Siso, C., Nance, J. D., Krummel, P. B., Mühle, J., Young,
D., O’Doherty, S., Salameh, P. K., Harth, C. M., Prinn, R. G.,
Weiss, R. F., et al. (2021). A decline in global CFC-11 emissions
during 2018–2019. Nature, 590(7846), 428–432. https://doi.org/
10.1038/s41586-021-03260-5
Solomon, S., Alcamo, J., & Ravishankara, A. R. (2020). Unfinished business after five decades of ozone-layer science and
policy. Nature Communications, 11(1), 4272. https://doi.org/10.
1038/s41467-020-18052-0
Dameris, M., Jöckel, P., & Nützel, M. (2019). Possible implications of enhanced chlorofluorocarbon-11 concentrations on
ozone. Atmospheric Chemistry and Physics, 19(22), 13759–
13771. https://doi.org/10.5194/acp-19-13759-2019
Fleming, E. L., Newman, P. A., Liang, Q., & Daniel, J. S. (2020).
The Impact of Continuing CFC-11 Emissions on Stratospheric
Ozone. Journal of Geophysical Research. https:// doi. org/ 10.
1029/2019jd031849
Keeble, J., Abraham, N. L., Archibald, A. T., Chipperfield, M. P.,
Dhomse, S., Griffiths, P. T., & Pyle, J. A. (2020). Modelling the
potential impacts of the recent, unexpected increase in CFC-11
emissions on total column ozone recovery. Atmospheric Chemistry and Physics, 20(12), 7153–7166. https://doi.org/10.5194/
acp-20-7153-2020
Caldwell, M. M. (1971). Solar UV irradiation and the growth
and development of higher plants. In A. C. Giese (Ed.), Current
Topics in Photobiology and Photochemistry, Photophysiology.
Academic Press.
Young, P. J., Harper, A. B., Huntingford, C., Paul, N. D., Morgenstern, O., Newman, P. A., Oman, L. D., Madronich, S., & Garcia,
R. R. (2021). The Montreal Protocol protects the terrestrial carbon sink. Nature, 596(7872), 384–388. https://doi.org/10.1038/
s41586-021-03737-3
Ballaré, C. L., Caldwell, M. M., Flint, S. D., Robinson, S. A., &
Bornman, J. F. (2011). Effects of solar ultraviolet radiation on
terrestrial ecosystems. Patterns, mechanisms, and interactions
with climate change. Photochemical & Photobiological Sciences.
https://doi.org/10.1039/C0PP90035D
Day, T. A. (1993). Relating UV-B radiation screening effectiveness of foliage to absorbing-compound concentration and
Photochemical & Photobiological Sciences (2023) 22:937–989
93.
94.
95.
96.
97.
98.
99.
100.
101.
102.
103.
104.
anatomical characteristics in a diverse group of plants. Oecologia, 95(4), 542–550. https://doi.org/10.1007/BF00317439
Jansen, M. A. K., Ač, A., Klem, K., & Urban, O. (2022). A
meta-analysis of the interactive effects of UV and drought on
plants. Plant, Cell & Environment, 45(1), 41–54. https://doi.org/
10.1111/pce.14221
Rozema, J., Björn, L. O., Bornman, J. F., Gaberščik, A., Häder,
D. P., Trošt, T., Germ, M., Klisch, M., Gröniger, A., & Sinha, R.
P. (2002). The role of UV-B radiation in aquatic and terrestrial
ecosystems—an experimental and functional analysis of the evolution of UV-absorbing compounds. Journal of Photochemistry
and Photobiology B: Biology, 66(1), 2–12. https://doi.org/10.
1016/S1011-1344(01)00269-X
Waterman, M. J., Nugraha, A. S., Hendra, R., Ball, G. E., Robinson, S. A., & Keller, P. A. (2017). Antarctic moss biflavonoids
show high antioxidant and ultraviolet-screening activity. Journal
of Natural Products, 80(8), 2224–2231. https://doi.org/10.1021/
acs.jnatprod.7b00085
Williamson, C. E., Zepp, R. G., Lucas, R. M., Madronich, S.,
Austin, A. T., Ballare, C. L., Norval, M., Sulzberger, B., Bais, A.
F., McKenzie, R. L., Robinson, S. A., Häder, D.-P., Paul, N. D.,
& Bornman, J. F. (2014). Solar ultraviolet radiation in a changing
climate. Nature Climate Change, 4(6), 434–441. https://doi.org/
10.1038/nclimate2225
Son, S.-W., Han, B.-R., Garfinkel, C. I., Kim, S.-Y., Park, R.,
Abraham, N. L., Akiyoshi, H., Archibald, A. T., Butchart, N.,
Chipperfield, M. P., Dameris, M., Deushi, M., Dhomse, S. S.,
Hardiman, S. C., Jöckel, P., Kinnison, D., Michou, M., Morgenstern, O., O’Connor, F. M., Oman, L. D., et al. (2018). Tropospheric jet response to Antarctic ozone depletion: An update with
Chemistry-Climate Model Initiative (CCMI) models. Environmental Research Letters. https:// doi. org/ 10. 1088/ 1748- 9326/
aabf21
Seviour, W. J. M., Codron, F., Doddridge, E. W., Ferreira, D.,
Gnanadesikan, A., Kelley, M., Kostov, Y., Marshall, J., Polvani,
L. M., Thomas, J. L., & Waugh, D. W. (2019). The Southern
Ocean sea surface temperature response to ozone depletion: A
multimodel comparison. Journal of Climate, 32(16), 5107–5121.
https://doi.org/10.1175/jcli-d-19-0109.1
Liu, W., Hegglin, M. I., Checa-Garcia, R., Li, S., Gillett, N. P.,
Lyu, K., Zhang, X., & Swart, N. C. (2022). Stratospheric ozone
depletion and tropospheric ozone increases drive Southern Ocean
interior warming. Nature Climate Change, 12(4), 365–372.
https://doi.org/10.1038/s41558-022-01320-w
Lu, X., Zhang, L., Zhao, Y., Jacob, D. J., Hu, Y., Hu, L., Gao, M.,
Liu, X., Petropavlovskikh, I., McClure-Begley, A., & Querel, R.
(2019). Surface and tropospheric ozone trends in the Southern
Hemisphere since 1990: Possible linkages to poleward expansion of the Hadley circulation. Science Bulletin, 64(6), 400–409.
https://doi.org/10.1016/j.scib.2018.12.021
Turner, J., & Comiso, J. (2017). Solve Antarctica’s sea-ice puzzle. Nature, 547, 275–277. https://doi.org/10.1038/547275a
Turner, J., Phillips, T., Marshall, G. J., Hosking, J. S., Pope, J. O.,
Bracegirdle, T. J., & Deb, P. (2017). Unprecedented springtime
retreat of Antarctic sea ice in 2016. Geophysical Research Letters, 44(13), 6868–6875. https://doi.org/10.1002/2017gl073656
Parkinson, C. L. (2019). A 40-y record reveals gradual Antarctic
sea ice increases followed by decreases at rates far exceeding the
rates seen in the Arctic. Proceedings of the National Academy of
Sciences, 116(29), 14414–14423. https://doi.org/10.1073/pnas.
1906556116
Gillett, Z. E., Arblaster, J. M., Dittus, A. J., Deushi, M., Jöckel, P.,
Kinnison, D. E., Morgenstern, O., Plummer, D. A., Revell, L. E.,
Rozanov, E., Schofield, R., Stenke, A., Stone, K. A., & Tilmes,
S. (2019). Evaluating the relationship between interannual variations in the Antarctic ozone hole and Southern Hemisphere
979
105.
106.
107.
108.
109.
110.
111.
112.
113.
114.
115.
116.
surface climate in chemistry–climate models. Journal of Climate,
32(11), 3131–3151. https://doi.org/10.1175/jcli-d-18-0273.1
Fogt, R. L., & Marshall, G. J. (2020). The Southern Annular
Mode: Variability, trends, and climate impacts across the Southern Hemisphere. WIREs Climate Change. https:// doi. org/ 10.
1002/wcc.652
Morales, M. S., Cook, E. R., Barichivich, J., Christie, D. A.,
Villalba, R., LeQuesne, C., Srur, A. M., Ferrero, M. E., Gonzalez-Reyes, A., Couvreux, F., Matskovsky, V., Aravena, J. C.,
Lara, A., Mundo, I. A., Rojas, F., Prieto, M. R., Smerdon, J.
E., Bianchi, L. O., Masiokas, M. H., Urrutia-Jalabert, R., et al.
(2020). Six hundred years of South American tree rings reveal an
increase in severe hydroclimatic events since mid-20th century.
Proceedings of the National Academy of Sciences of the United
States of America, 117(29), 16816–16823. https://doi.org/10.
1073/pnas.2002411117
Damiani, A., Cordero, R. R., Llanillo, P. J., Feron, S., Boisier, J.
P., Garreaud, R., Rondanelli, R., Irie, H., & Watanabe, S. (2020).
Connection between Antarctic ozone and climate: interannual
precipitation changes in the Southern Hemisphere. Atmosphere.
https://doi.org/10.3390/atmos11060579
Banerjee, A., Fyfe, J. C., Polvani, L. M., Waugh, D., & Chang,
K.-L. (2020). A pause in Southern Hemisphere circulation trends
due to the Montreal Protocol. Nature, 579(7800), 544–548.
https://doi.org/10.1038/s41586-020-2120-4
Bracegirdle, T. J., Krinner, G., Tonelli, M., Haumann, F. A.,
Naughten, K. A., Rackow, T., Roach, L. A., & Wainer, I. (2020).
Twenty first century changes in Antarctic and Southern Ocean
surface climate in CMIP6. Atmospheric Science Letters. https://
doi.org/10.1002/asl.984
Lim, E.-P., Hendon, H. H., Boschat, G., Hudson, D., Thompson,
D. W. J., Dowdy, A. J., & Arblaster, J. M. (2019). Australian hot
and dry extremes induced by weakenings of the stratospheric
polar vortex. Nature Geoscience, 12(11), 896–901. https://doi.
org/10.1038/s41561-019-0456-x
Hendon, H. H., Thompson, D. J. W., Lim, E.-P., Butler, A. H.,
Newman, P. A., Coy, L., Scaife, A., Polichtchouk, I., Garreaud,
R. S., & T.G., S., & Nakamura, H. (2019). Rare forecasted climate event under way in the Southern Hemisphere. Nature,
573(7775), 495. https://doi.org/10.1038/d41586-019-02858-0
Lim, E.-P., Hendon, H. H., Butler, A. H., Garreaud, R. D., Polichtchouk, I., Shepherd, T. G., Scaife, A., Comer, R., Coy, L.,
Newman, P. A., Thompson, D. J. W., & Nakamura, H. (2020).
The 2019 Antarctic sudden stratospheric warming. SPARC Newsletter, 54, 10–13.
Noguchi, S., Kuroda, Y., Kodera, K., & Watanabe, S. (2020).
Robust enhancement of tropical convective activity by the 2019
Antarctic sudden stratospheric warming. Geophysical Research
Letters. https://doi.org/10.1029/2020GL088743
Robinson, S. A., Klekociuk, A. R., King, D. H., Pizarro Rojas,
M., Zúñiga, G. E., & Bergstrom, D. M. (2020). The 2019/2020
summer of Antarctic heatwaves. Global Change Biology, 26(6),
3178–3180. https://doi.org/10.1111/gcb.15083
Newman, P., Nash, E. R., Kramarova, N., & Butler, A. (2020).
The 2019 southern stratospheric warming. In State of the Climate
in 2019, Bulletin of the American Meteorological Society, 101(8),
S297–S298. https://doi.org/10.1175/BAMS-D-20-0090.1
Lim, E.-P., Hendon, H. H., Butler, A. H., Thompson, D. W. J.,
Lawrence, Z., Scaife, A. A., Shepherd, T. G., Polichtchouk,
I., Nakamura, H., Kobayashi, C., Comer, R., Coy, L., Dowdy,
A., Garreaud, R. D., Newman, P. A., & Wang, G. (2021). The
2019 Southern Hemisphere stratospheric polar vortex weakening and its impacts. Bulletin of the American Meteorological Society, 102(6), E1150–E1171. https:// doi. org/ 10. 1175/
BAMS-D-20-0112.1
13
980
117. Jucker, M., & Goyal, R. (2022). Ozone-Forced Southern Annular
Mode During Antarctic Stratospheric Warming Events. Geophysical Research Letters. https://doi.org/10.1029/2021gl095270
118. Jucker, M., Reichler, T., & Waugh, D. W. (2021). How frequent
are Antarctic sudden stratospheric warmings in present and
future climate? Geophysical Research Letters. https://doi.org/
10.1029/2021GL093215
119. Kablick, G. P., III., Allen, D. R., Fromm, M. D., & Nedoluha,
G. E. (2020). Australian PyroCb smoke generates synoptic-scale
stratospheric anticyclones. Geophysical Research Letters. https://
doi.org/10.1029/2020GL088101
120. Ohneiser, K., Ansmann, A., Kaifler, B., Chudnovsky, A., Barja,
B., Knopf, D. A., Kaifler, N., Baars, H., Seifert, P., Villanueva,
D., Jimenez, C., Radenz, M., Engelmann, R., Veselovskii, I., &
Zamorano, F. (2022). Australian wildfire smoke in the stratosphere: The decay phase in 2020/2021 and impact on ozone
depletion. Atmospheric Chemistry and Physics, 22(11), 7417–
7442. https://doi.org/10.5194/acp-22-7417-2022
121. Ohneiser, K., Ansmann, A., Baars, H., Seifert, P., Barja, B., Jimenez, C., Radenz, M., Teisseire, A., Floutsi, A., Haarig, M., Foth,
A., Chudnovsky, A., Engelmann, R., Zamorano, F., Bühl, J., &
Wandinger, U. (2020). Smoke of extreme Australian bushfires
observed in the stratosphere over Punta Arenas, Chile, in January
2020: Optical thickness, lidar ratios, and depolarization ratios at
355 and 532 nm. Atmospheric Chemistry and Physics, 20(13),
8003–8015. https://doi.org/10.5194/acp-20-8003-2020
122. Boone, C. D., Bernath, P. F., & Fromm, M. D. (2020). Pyrocumulonimbus stratospheric plume injections measured by the ACEFTS. Geophysical Research Letters, 47(15), e2020GL088442.
https://doi.org/10.1029/2020GL088442
123. Khaykin, S., Legras, B., Bucci, S., Sellitto, P., Isaksen, L., Tencé,
F., Bekki, S., Bourassa, A., Rieger, L., Zawada, D., Jumelet, J.,
& Godin-Beekmann, S. (2020). The 2019/20 Australian wildfires generated a persistent smoke-charged vortex rising up to 35
km altitude. Communications Earth & Environment, 1(1), 1–12.
https://doi.org/10.1038/s43247-020-00022-5
124. Hirsch, E., & Koren, I. (2021). Record-breaking aerosol levels
explained by smoke injection into the stratosphere. Science,
371(6535), 1269–1274. https://doi.org/10.1126/science.abe1415
125. Allen, D. R., Fromm, M. D., Kablick, G. P., III., & Nedoluha, G.
E. (2020). Smoke with Induced Rotation and Lofting (SWIRL)
in the stratosphere. Journal of the Atmospheric Sciences, 77(12),
4297–4316. https://doi.org/10.1175/JAS-D-20-0131.1
126. Yu, P., Davis, S. M., Toon, O. B., Portmann, R. W., Bardeen,
C. G., Barnes, J. E., Telg, H., Maloney, C., & Rosenlof, K. H.
(2021). Persistent stratospheric warming due to 2019–2020 Australian wildfire smoke. Geophysical Research Letters. https://doi.
org/10.1029/2021GL092609
127. Schwartz, M. J., Santee, M. L., Pumphrey, H. C., Manney, G. L.,
Lambert, A., Livesey, N. J., Millán, L., Neu, J. L., Read, W. G.,
& Werner, F. (2020). Australian new year’s pyroCb impact on
stratospheric composition. Geophysical Research Letters. https://
doi.org/10.1029/2020GL090831
128. Anderson, J. G., Wilmouth, D. M., Smith, J. B., & Sayres, D. S.
(2012). UV dosage levels in summer: Increased risk of ozone
loss from convectively injected water vapor. Science, 337(6096),
835–839. https://doi.org/10.1126/science.1222978
129. Schwartz, M. J., Read, W. G., Santee, M. L., Livesey, N. J.,
Froidevaux, L., Lambert, A., & Manney, G. L. (2013). Convectively injected water vapor in the North American summer
lowermost stratosphere. Geophysical Research Letters, 40(10),
2316–2321. https://doi.org/10.1002/grl.50421
130. Bernath, P., Boone, C., & Crouse, J. (2022). Wildfire smoke
destroys stratospheric ozone. Science, 375(6586), 1292–1295.
https://doi.org/10.1126/science.abm5611
13
Photochemical & Photobiological Sciences (2023) 22:937–989
131. Solomon, S., Dube, K., Stone, K., Yu, P., Kinnison, D., Toon,
O. B., Strahan, S. E., Rosenlof, K. H., Portmann, R., Davis, S.,
Randel, W., Bernath, P., Boone, C., Bardeen, C. G., Bourassa, A.,
Zawada, D., & Degenstein, D. (2022). On the stratospheric chemistry of midlatitude wildfire smoke. Proceedings of the National
Academy of Sciences. https://doi.org/10.1073/pnas.2117325119
132. Santee, M. L., Lambert, A., Manney, G. L., Livesey, N. J.,
Froidevaux, L., Neu, J. L., Schwartz, M. J., Millán, L. F., Werner,
F., Read, W. G., Park, M., Fuller, R. A., & Ward, B. M. (2022).
Prolonged and pervasive perturbations in the composition of
the Southern Hemisphere midlatitude lower stratosphere from
the Australian New Year’s fires. Geophysical Research Letters.
https://doi.org/10.1029/2021GL096270
133. Polvani, L. M., Banerjee, A., Chemke, R., Doddridge, E. W.,
Ferreira, D., Gnanadesikan, A., Holland, M. A., Kostov, Y., Marshall, J., Seviour, W. J. M., Solomon, S., & Waugh, D. W. (2021).
Interannual SAM modulation of Antarctic sea ice extent does not
account for its long-term trends, pointing to a limited role for
ozone depletion. Geophysical Research Letters. https://doi.org/
10.1029/2021gl094871
134. Haase, S., & Matthes, K. (2019). The importance of interactive
chemistry for stratosphere–troposphere coupling. Atmospheric
Chemistry and Physics, 19(5), 3417–3432. https://doi.org/10.
5194/acp-19-3417-2019
135. Zhou, C., Zhang, T., & Zheng, L. (2019). The characteristics of
surface albedo change trends over the Antarctic sea ice region
during recent decades. Remote Sensing. https://doi.org/10.3390/
rs11070821
136. Meehl, G. A., Arblaster, J. M., Chung, C. T. Y., Holland, M. M.,
DuVivier, A., Thompson, L., Yang, D., & Bitz, C. M. (2019).
Sustained ocean changes contributed to sudden Antarctic sea ice
retreat in late 2016. Nature Communications, 10(1), 14. https://
doi.org/10.1038/s41467-018-07865-9
137. Wang, G., Hendon, H. H., Arblaster, J. M., Lim, E.-P., Abhik,
S., & van Rensch, P. (2019). Compounding tropical and stratospheric forcing of the record low Antarctic sea-ice in 2016.
Nature Communications, 10(1), 13. https:// doi. org/ 10. 1038/
s41467-018-07689-7
138. Wang, Z., Turner, J., Wu, Y., & Liu, C. (2019). Rapid decline of
total Antarctic sea ice extent during 2014–16 controlled by winddriven sea ice drift. Journal of Climate, 32(17), 5381–5395.
https://doi.org/10.1175/jcli-d-18-0635.1
139. Doddridge, E. W., & Marshall, J. (2017). Modulation of the seasonal cycle of Antarctic sea ice extent related to the southern
annular mode. Geophysical Research Letters, 44(19), 9761–
9768. https://doi.org/10.1002/2017gl074319
140. Ferreira, D., Marshall, J., Bitz, C. M., Solomon, S., & Plumb,
A. (2015). Antarctic ocean and sea ice response to ozone depletion: A two-time-scale problem. Journal of Climate, 28(3),
1206–1226. https://doi.org/10.1175/jcli-d-14-00313.1
141. Seviour, W. J. M., Codron, F., Doddridge, E. W., Ferreira,
D., Gnanadesikan, A., Kelley, M., Kostov, Y., Marshall, J., Polvani, L. M., Thomas, J. L., & Waugh, D. W. The Southern
Ocean sea surface temperature response to ozone depletion:
A multimodel comparison. Journal of Climate, 32(16), 5107–
5121. https://doi.org/10.1175/jcli-d-19-0109.1
142. Xia, Y., Hu, Y., Liu, J., Huang, Y., Xie, F., & Lin, J. (2020).
Stratospheric ozone-induced cloud radiative effects on Antarctic sea ice. Advances in Atmospheric Sciences, 37(5), 505–514.
https://doi.org/10.1007/s00376-019-8251-6
143. Li, S., Liu, W., Lyu, K., & Zhang, X. (2021). The effects of
historical ozone changes on Southern Ocean heat uptake and
storage. Climate Dynamics, 57(7–8), 2269–2285. https://doi.org/
10.1007/s00382-021-05803-y
144. England, M., Polvani, L., & Sun, L. (2018). Contrasting the
Antarctic and Arctic atmospheric responses to projected sea ice
Photochemical & Photobiological Sciences (2023) 22:937–989
145.
146.
147.
148.
149.
150.
151.
152.
153.
154.
155.
156.
157.
158.
159.
loss in the late twenty-first century. Journal of Climate, 31(16),
6353–6370. https://doi.org/10.1175/jcli-d-17-0666.1
Doddridge, E. W., Marshall, J., Song, H., Campin, J.-M., &
Kelley, M. (2021). Southern Ocean heat storage, reemergence,
and winter sea ice decline induced by summertime winds.
Journal of Climate, 34(4), 1403–1415. https://doi.org/10.1175/
jcli-d-20-0322.1
Jakobs, C. L., Reijmer, C. H., van den Broeke, M. R., van de
Berg, W. J., & van Wessem, J. M. (2021). Spatial variability of
the snowmelt-albedo feedback in Antarctica. Journal of Geophysical Research. https://doi.org/10.1029/2020jf005696
Bergstrom, A., Gooseff, M. N., Myers, M., Doran, P. T., &
Cross, J. M. (2020). The seasonal evolution of albedo across
glaciers and the surrounding landscape of Taylor Valley. Antarctica. The Cryosphere, 14(3), 769–788. https://doi.org/10.5194/
tc-14-769-2020
Schneider, D. P., Kay, J. E., & Lenaerts, J. (2020). Improved
clouds over Southern Ocean amplify Antarctic precipitation
response to ozone depletion in an earth system model. Climate Dynamics, 55(5–6), 1665–1684. https://doi.org/10.1007/
s00382-020-05346-8
Lenaerts, J. T. M., Fyke, J., & Medley, B. (2018). The signature
of ozone depletion in recent Antarctic precipitation change: a
study with the Community Earth System Model. Geophysical
Research Letters. https://doi.org/10.1029/2018gl078608
Friedel, M., Chiodo, G., Stenke, A., Domeisen, D. I. V., Fueglistaler, S., Anet, J. G., & Peter, T. (2022). Springtime arctic
ozone depletion forces northern hemisphere climate anomalies.
Nature Geoscience, 15(7), 541–547. https:// doi. org/ 10. 1038/
s41561-022-00974-7
Domeisen, D. I. V., & Butler, A. H. (2020). Stratospheric
drivers of extreme events at the Earth’s surface. Communications Earth & Environment, 1(1), 1–8. https://doi.org/10.1038/
s43247-020-00060-z
Xia, Y., Hu, Y., Huang, Y., Zhao, C., Xie, F., & Yang, Y. (2021).
Significant contribution of severe ozone loss to the SiberianArctic surface warming in spring 2020. Geophysical Research
Letters. https://doi.org/10.1029/2021GL092509
Overland, J. E., & Wang, M. (2021). The 2020 Siberian heat
wave. International Journal of Climatology, 41, E2341–E2346.
https://doi.org/10.1002/joc.6850
Zhang, J., Tian, W., Pyle, J. A., Keeble, J., Abraham, N. L., Chipperfield, M. P., Xie, F., Yang, Q., Mu, L., Ren, H.-L., Wang, L.,
& Xu, M. (2022). Responses of Arctic sea ice to stratospheric
ozone depletion. Science Bulletin, 67(11), 1182–1190. https://
doi.org/10.1016/j.scib.2022.03.015
Baldwin, M. P., Ayarzagüena, B., Birner, T., Butchart, N., Butler,
A. H., Charlton-Perez, A. J., Domeisen, D. I. V., Garfinkel, C. I.,
Garny, H., Gerber, E. P., Hegglin, M. I., Langematz, U., & Pedatella, N. M. (2021). Sudden Stratospheric Warmings. Reviews of
Geophysics. https://doi.org/10.1029/2020rg000708
Butler, A. H., & Lee, S. H. (2022). The 2020 Arctic Sudden
Stratospheric Warming. In State of the Climate in 2021, Bulletin
of the American Meteorological Society, 103(8), S296–S298.
https://doi.org/10.1175/BAMS-D-22-0082.1
Huang, J., Hitchcock, P., Maycock, A. C., McKenna, C. M., &
Tian, W. (2021). Northern hemisphere cold air outbreaks are
more likely to be severe during weak polar vortex conditions.
Communications Earth & Environment. https://doi.org/10.1038/
s43247-021-00215-6
Zhang, M., Yang, X. Y., & Huang, Y. (2022). Impacts of sudden
stratospheric warming on extreme cold events in early 2021: an
ensemble-based sensitivity analysis. Geophysical Research Letters. https://doi.org/10.1029/2021GL096840
Xie, F., Ma, X., Li, J., Huang, J., Tian, W., Zhang, J., Hu, Y., Sun,
C., Zhou, X., Feng, J., & Yang, Y. (2018). An advanced impact
981
160.
161.
162.
163.
164.
165.
166.
167.
168.
169.
170.
171.
of Arctic stratospheric ozone changes on spring precipitation in
China. Climate Dynamics, 51(11–12), 4029–4041. https://doi.
org/10.1007/s00382-018-4402-1
Stone, K. A., Solomon, S., Kinnison, D. E., Baggett, C. F.,
& Barnes, E. A. (2019). Prediction of Northern Hemisphere
regional surface temperatures using stratospheric ozone information. Journal of Geophysical Research: Atmospheres, 124(12),
5922–5933. https://doi.org/10.1029/2018jd029626
Ma, X., Xie, F., Li, J., Zheng, X., Tian, W., Ding, R., Sun, C., &
Zhang, J. (2019). Effects of Arctic stratospheric ozone changes
on spring precipitation in the northwestern United States. Atmospheric Chemistry and Physics, 19(2), 861–875. https://doi.org/
10.5194/acp-19-861-2019
Bais, A. F., McKenzie, R. L., Bernhard, G., Aucamp, P. J., Ilyas,
M., Madronich, S., & Tourpali, K. (2015). Ozone depletion and
climate change: Impacts on UV radiation. Photochemical &
Photobiological Sciences, 14(1), 19–52. https://doi.org/10.1039/
c4pp90032d
McKenzie, R. L., Aucamp, P. J., Bais, A. F., Björn, L. O., Ilyas,
M., & Madronich, S. (2011). Ozone depletion and climate
change: Impacts on UV radiation. Photochemical & Photobiological Sciences, 10(2), 182–198. https://doi.org/10.1039/c0pp9
0034f
Campanelli, M., Diémoz, H., Siani, A. M., di Sarra, A., Iannarelli, A. M., Kudo, R., Fasano, G., Casasanta, G., Tofful, L.,
Cacciani, M., Sanò, P., & Dietrich, S. (2022). Aerosol optical
characteristics in the urban area of Rome, Italy, and their impact
on the UV index. Atmospheric Measurement Techniques, 15(5),
1171–1183. https://doi.org/10.5194/amt-15-1171-2022
Mok, J., Krotkov, N. A., Arola, A., Torres, O., Jethva, H.,
Andrade, M., Labow, G., Eck, T. F., Li, Z., & Dickerson, R.
R. (2016). Impacts of brown carbon from biomass burning on
surface UV and ozone photochemistry in the Amazon Basin.
Scientific Reports, 6, 36940. https://doi.org/10.1038/srep36940
Carlund, T., Kouremeti, N., Kazadzis, S., & Gröbner, J. (2017).
Aerosol optical depth determination in the UV using a fourchannel precision filter radiometer. Atmospheric Measurement Techniques, 10(3), 905–923. https:// doi. org/ 10. 5194/
amt-10-905-2017
López-Solano, J., Redondas, A., Carlund, T., RodriguezFranco, J. J., Diémoz, H., León-Luis, S. F., Hernández-Cruz, B.,
Guirado-Fuentes, C., Kouremeti, N., Gröbner, J., Kazadzis, S.,
Carreño, V., Berjón, A., Santana-Díaz, D., Rodríguez-Valido,
M., De Bock, V., Moreta, J. R., Rimmer, J., Smedley, A. R. D.,
Boulkelia, L., et al. (2018). Aerosol optical depth in the European
Brewer Network. Atmospheric Chemistry and Physics, 18(6),
3885–3902. https://doi.org/10.5194/acp-18-3885-2018
Mok, J., Krotkov, N. A., Torres, O., Jethva, H., Li, Z., Kim, J.,
Koo, J.-H., Go, S., Irie, H., Labow, G., Eck, T. F., Holben, B. N.,
Herman, J., Loughman, R. P., Spinei, E., Lee, S. S., Khatri, P., &
Campanelli, M. (2018). Comparisons of spectral aerosol single
scattering albedo in Seoul. South Korea. Atmospheric Measurement Techniques, 11(4), 2295–2311. https:// doi. org/ 10. 5194/
amt-11-2295-2018
Fountoulakis, I., Natsis, A., Siomos, N., Drosoglou, T., & Bais,
A. F. (2019). Deriving aerosol absorption properties from solar
ultraviolet radiation spectral measurements at Thessaloniki.
Greece. Remote Sensing, 11(18), 2179. https://doi.org/10.3390/
rs11182179
Rimmer, J. S., Redondas, A., & Karppinen, T. (2018).
EuBrewNet–A European brewer network (COST Action
ES1207), an overview. Atmospheric Chemistry and Physics, 18(14), 10347–10353. https:// doi. org/ 10. 5194/
acp-18-10347-2018
Ipiña, A., López-Padilla, G., Retama, A., Piacentini, R. D., &
Madronich, S. (2021). Ultraviolet radiation environment of a
13
982
172.
173.
174.
175.
176.
177.
178.
179.
180.
181.
182.
183.
184.
Photochemical & Photobiological Sciences (2023) 22:937–989
tropical megacity in transition: Mexico City 2000–2019. Environmental Science & Technology, 55(16), 10946–10956. https://
doi.org/10.1021/acs.est.0c08515
Wilson, S. R., Madronich, S., Longstreth, J. D., & Solomon, K.
R. (2019). Interactive effects of changing stratospheric ozone
and climate on tropospheric composition and air quality, and the
consequences for human and ecosystem health. Photochemical
& Photobiological Sciences, 18(3), 775–803. https://doi.org/10.
1039/c8pp90064g
Moses, E., Cardenas, B., Nagpure, A., & Pai, M. (2020). Can
An Airshed Governance Framework in India Spur Clean Air for
All? Lessons from Mexico City and Los Angeles: Policy Brief,
CCAPC/2020/01. Collaborative Clean Air Policy Centre, India.
Cabrera, S., Ipiña, A., Damiani, A., Cordero, R. R., & Piacentini,
R. D. (2012). UV index values and trends in Santiago, Chile
(33.5°S) based on ground and satellite data. Journal of Photochemistry & Photobiology B: Biology, 115, 73–84. https://doi.
org/10.1016/j.jphotobiol.2012.06.013
Roshan, D. R., Koc, M., Abdallah, A., Martin-Pomares, L.,
Isaifan, R., & Fountoukis, C. (2020). UV Index forecasting
under the influence of desert dust: evaluation against surface
and satellite-retrieved data. Atmosphere. https://doi.org/10.3390/
atmos11010096
du Preez, D. J., Bencherif, H., Portafaix, T., Lamy, K., & Wright,
C. Y. (2021). Solar ultraviolet radiation in Pretoria and its relations to aerosols and tropospheric ozone during the biomass
burning season. Atmosphere. https://doi.org/10.3390/atmos12020
132
Diffey, B. L., Jansén, C. T., Urbach, F., & Wulf, H. C. (1997).
The standard erythema dose: A new photobiological concept.
Photodermatology, Photoimmunology and Photomedicine, 13(1–
2), 64–66. https://doi.org/10.1111/j.1600-0781.1997.tb00110.x
Fitzpatrick, T. B. (1988). The Validity and Practicality of SunReactive Skin Types I Through VI. Archives of Dermatology,
124(6), 869–871. https://doi.org/10.1001/archderm.1988.01670
060015008
McKenzie, R. L., Liley, J. B., & Björn, L. O. (2009). UV Radiation: Balancing Risks and Benefits. Photochemistry and Photobiology, 85, 88–98. https://doi.org/10.1111/j.1751-1097.2008.
00400.x
Igoe, D. P., Parisi, A. V., Downs, N. J., & Butler, H. (2022). A
case study of UV exposure risk in Sydney during the 2019/2020
New South Wales bushfires. Photochemistry and Photobiology,
98(5), 1236–1244. https://doi.org/10.1111/php.13603
ICNIRP. (2010). ICNIRP Statement—Protection of workers
against ultraviolet radiation. Health Physics, 99(1), 66–87.
https://doi.org/10.1097/HP.0b013e3181d85908
Mortier, A., Gliß, J., Schulz, M., Aas, W., Andrews, E., Bian, H.,
Chin, M., Ginoux, P., Hand, J., Holben, B., Zhang, H., Kipling,
Z., Kirkevåg, A., Laj, P., Lurton, T., Myhre, G., Neubauer, D.,
Olivié, D., von Salzen, K., Skeie, R. B., et al. (2020). Evaluation of climate model aerosol trends with ground-based observations over the last 2 decades – an AeroCom and CMIP6 analysis. Atmospheric Chemistry and Physics, 20(21), 13355–13378.
https://doi.org/10.5194/acp-20-13355-2020
Tong, D., Cheng, J., Liu, Y., Yu, S., Yan, L., Hong, C., Qin,
Y., Zhao, H., Zheng, Y., Geng, G., Li, M., Liu, F., Zhang, Y.,
Zheng, B., Clarke, L., & Zhang, Q. (2020). Dynamic projection of anthropogenic emissions in China: Methodology and
2015–2050 emission pathways under a range of socio-economic,
climate policy, and pollution control scenarios. Atmospheric
Chemistry and Physics, 20(9), 5729–5757. https://doi.org/10.
5194/acp-20-5729-2020
Drugé, T., Nabat, P., Mallet, M., & Somot, S. (2021). Future
evolution of aerosols and implications for climate change in
the Euro-Mediterranean region using the CNRM-ALADIN63
13
185.
186.
187.
188.
189.
190.
191.
192.
193.
194.
195.
196.
197.
198.
regional climate model. Atmospheric Chemistry and Physics,
21(10), 7639–7669. https://doi.org/10.5194/acp-21-7639-2021
Millán, L. F., & Manney, G. L. (2017). An assessment of ozone
mini-hole representation in reanalyses over the Northern Hemisphere. Atmospheric Chemistry and Physics, 17(15), 9277–9289.
https://doi.org/10.5194/acp-17-9277-2017
Raptis, I.-P., Eleftheratos, K., Kazadzis, S., Kosmopoulos,
P., Papachristopoulou, K., & Solomos, S. (2021). The combined effect of ozone and aerosols on erythemal irradiance in
an extremely low ozone event during May 2020. Atmosphere.
https://doi.org/10.3390/atmos12020145
Becherini, F., Vitale, V., Lupi, A., Stone, R. S., Salvatori, R.,
Salzano, R., di Carlo, P., Viola, A. P., & Mazzola, M. (2021).
Surface albedo and spring snow melt variations at Ny-Ålesund,
Svalbard, Bulletin of Atmospheric Science and Technology.
https://doi.org/10.1007/s42865-021-00043-8
Brunt, K. M., Sergienko, O., & MacAyeal, D. R. (2017). Observations of unusual fast-ice conditions in the southwest Ross
Sea, Antarctica: Preliminary analysis of iceberg and storminess
effects. Annals of Glaciology, 44, 183–187. https://doi.org/10.
3189/172756406781811754
Kim, S., Saenz, B., Scanniello, J., Daly, K., & Ainley, D. (2018).
Local climatology of fast ice in McMurdo Sound. Antarctica.
Antarctic Science, 30(2), 125–142. https://doi.org/10.1017/s0954
102017000578
Bernhard, G., & Stierle, S. (2020). Trends of UV radiation in
Antarctica. Atmosphere, 11(8), 795. https:// doi. org/ 10. 3390/
atmos11080795
Gordon, E. M., Seppälä, A., Funke, B., Tamminen, J., & Walker,
K. A. (2021). Observational evidence of energetic particle precipitation NOx (EPP-NOx) interaction with chlorine curbing
Antarctic ozone loss. Atmospheric Chemistry and Physics, 21(4),
2819–2836. https://doi.org/10.5194/acp-21-2819-2021
Gordon, E. M., Seppälä, A., & Tamminen, J. (2020). Evidence
for energetic particle precipitation and quasi-biennial oscillation modulations of the Antarctic NO2 springtime stratospheric column from OMI observations. Atmospheric Chemistry and Physics, 20(11), 6259–6271. https://doi.org/10.5194/
acp-20-6259-2020
Haigh, J. D., Winning, A. R., Toumi, R., & Harder, J. W. (2010).
An influence of solar spectral variations on radiative forcing of
climate. Nature, 467(7316), 696–699. https://doi.org/10.1038/
nature09426
Xiao, Z.-N., Dong, S., & Zhong, Q. (2019). Numerical simulation
of climate response to ultraviolet irradiation forcing. Advances
in Climate Change Research, 10(3), 133–142. https://doi.org/10.
1016/j.accre.2019.07.001
Roy, I., & Haigh, J. D. (2011). The influence of solar variability and the quasi-biennial oscillation on lower atmospheric
temperatures and sea level pressure. Atmospheric Chemistry
and Physics, 11(22), 11679–11687. https:// doi. org/ 10. 5194/
acp-11-11679-2011
Arsenovic, P., Rozanov, E., Anet, J., Stenke, A., Schmutz, W.,
& Peter, T. (2018). Implications of potential future grand solar
minimum for ozone layer and climate. Atmospheric Chemistry and Physics, 18(5), 3469–3483. https:// doi. org/ 10. 5194/
acp-18-3469-2018
Miyahara, H., Tokanai, F., Moriya, T., Takeyama, M., Sakurai,
H., Horiuchi, K., & Hotta, H. (2021). Gradual onset of the Maunder Minimum revealed by high-precision carbon-14 analyses.
Scientific Reports. https://doi.org/10.1038/s41598-021-84830-5
Ineson, S., Maycock, A. C., Gray, L. J., Scaife, A. A., Dunstone,
N. J., Harder, J. W., Knight, J. R., Lockwood, M., Manners, J. C.,
& Wood, R. A. (2015). Regional climate impacts of a possible
future grand solar minimum. Nature Communications. https://
doi.org/10.1038/ncomms8535
Photochemical & Photobiological Sciences (2023) 22:937–989
199. Kitiashvili, I. N. (2020). Application of synoptic magnetograms
to global solar activity forecast. The Astrophysical Journal.
https://doi.org/10.3847/1538-4357/ab64e7
200. Zharkova, V. (2020). Modern grand solar minimum will lead to
terrestrial cooling. Temperature, 7(3), 217–222. https://doi.org/
10.1080/23328940.2020.1796243
201. Stone, K. A., Solomon, S., Kinnison, D. E., Pitts, M. C., Poole,
L. R., Mills, M. J., Schmidt, A., Neely, R. R., Ivy, D., Schwartz,
M. J., Vernier, J.-P., Johnson, B. J., Tully, M. B., Klekociuk, A.
R., König-Langlo, G., & Hagiya, S. (2017). Observing the impact
of Calbuco volcanic aerosols on south polar ozone depletion in
2015. Journal of Geophysical Research: Atmospheres, 122(21),
11862–11879. https://doi.org/10.1002/2017JD026987
202. Arnold, F., Bührke, T., & Qiu, S. (1990). Evidence for stratospheric ozone-depleting heterogeneous chemistry on volcanic
aerosols from El Chichón. Nature, 348(6296), 49–50. https://doi.
org/10.1038/348049a0
203. Brasseur, G., & Granier, C. (1992). Mount Pinatubo aerosols,
chlorofluorocarbons, and ozone depletion. Science, 257(5074),
1239–1242. https://doi.org/10.1126/science.257.5074.1239
204. Osipov, S., Stenchikov, G., Tsigaridis, K., LeGrande, A. N.,
Bauer, S. E., Fnais, M., & Lelieveld, J. (2021). The Toba supervolcano eruption caused severe tropical stratospheric ozone
depletion. Communications Earth & Environment. https://doi.
org/10.1038/s43247-021-00141-7
205. Labitzke, K., & McCormick, M. P. (1992). Stratospheric temperature increases due to Pinatubo aerosols. Geophysical Research
Letters, 19(2), 207–210. https://doi.org/10.1029/91GL02940
206. Brenna, H., Kutterolf, S., & Krüger, K. (2019). Global ozone
depletion and increase of UV radiation caused by pre-industrial
tropical volcanic eruptions. Scientific Reports. https://doi.org/10.
1038/s41598-019-45630-0
207. Ming, A., Winton, V. H. L., Keeble, J., Abraham, N. L., Dalvi,
M. C., Griffiths, P., Caillon, N., Jones, A. E., Mulvaney, R., Savarino, J., Frey, M. M., & Yang, X. (2020). Stratospheric ozone
changes from explosive tropical volcanoes: modeling and ice
core constraints. Journal of Geophysical Research. https://doi.
org/10.1029/2019JD032290
208. Krüger, K., Kutterolf, S., Hansteen, T. H., Schmidt, A., Fristad,
K. E., & Elkins-Tanton, L. T. (2015). Halogen release from Plinian eruptions and depletion of stratospheric ozone. Volcanism
and Global Environmental Change (pp. 244–259). Cambridge
University Press.
209. Mankin, W. G., Coffey, M. T., & Goldman, A. (1992). Airborne
observations of SO2, HCl, and O3 in the stratospheric plume of
the Pinatubo volcano in July 1991. Geophysical Research Letters,
19(2), 179–182. https://doi.org/10.1029/91gl02942
210. Zaratti, F., Piacentini, R. D., Guillén, H. A., Cabrera, S. H., Liley,
J. B., & McKenzie, R. L. (2014). Proposal for a modification of
the UVI risk scale. Photochemical & Photobiological Sciences,
13(7), 980–985. https://doi.org/10.1039/C4PP00006D
211. Norris, J. R., Allen, R. J., Evan, A. T., Zelinka, M. D., O’Dell,
C. W., & Klein, S. A. (2016). Evidence for climate change in the
satellite cloud record. Nature, 536(7614), 72–75. https://doi.org/
10.1038/nature18273
212. Bodas-Salcedo, A., Hill, P. G., Furtado, K., Williams, K. D.,
Field, P. R., Manners, J. C., Hyder, P., & Kato, S. (2016). Large
contribution of supercooled liquid clouds to the solar radiation
budget of the Southern Ocean. Journal of Climate, 29(11), 4213–
4228. https://doi.org/10.1175/JCLI-D-15-0564.1
213. Cherian, R., & Quaas, J. (2020). Trends in AOD, clouds, and
cloud radiative effects in satellite data and CMIP5 and CMIP6
model simulations over aerosol source regions. Geophysical
Research Letters. https://doi.org/10.1029/2020gl087132
214. Wu, D. L., Lee, J. N., Kim, K.-M., & Lim, Y.-K. (2020). Interannual variations of TOA albedo over the Arctic, Antarctic and
983
215.
216.
217.
218.
219.
220.
221.
222.
223.
224.
225.
226.
Tibetan Plateau in 2000–2019. Remote Sensing. https://doi.org/
10.3390/rs12091460
Bernhard, G. H., Fioletov, V. E., Grooß, J.-U., Ialongo, I., Johnsen, B., Lakkala, K., Manney, G. L., & Müller, R. (2019). Ozone
and UV radiation. In State of the Climate in 2018, Bulletin of the
American Meteorological Society. https://doi.org/10.1175/2019B
AMSStateoftheClimate.1
Rösner, B., Benedict, I., Van Heerwaarden, C., Weerts, A., Hazeleger, W., Bissolli, P., & Trachte, K. (2019). The long heat wave
and drought in Europe in 2018. In State of the Climate in 2018,
Bulletin of the American Meteorological Society. https://doi.org/
10.1175/2019BAMSStateoftheClimate.1
Marín, J. C., Bozkurt, D., & Barrett, B. S. (2022). Atmospheric
blocking trends and seasonality around the Antarctic Peninsula.
Journal of Climate, 35(12), 3803–3818. https://doi.org/10.1175/
jcli-d-21-0323.1
Wachowicz, L. J., Preece, J. R., Mote, T. L., Barrett, B. S., &
Henderson, G. R. (2020). Historical trends of seasonal Greenland
blocking under different blocking metrics. International Journal
of Climatology. https://doi.org/10.1002/joc.6923
Woollings, T., Barriopedro, D., Methven, J., Son, S.-W., Martius, O., Harvey, B., Sillmann, J., Lupo, A. R., & Seneviratne,
S. (2018). Blocking and its response to climate change. Current
Climate Change Reports, 4(3), 287–300. https://doi.org/10.1007/
s40641-018-0108-z
Dobson, G. M. B., Brewer, A. W., & Cwilong, B. M. (1946).
Bakerian Lecture: Meteorology of the lower stratosphere. Proceedings of the Royal Society of London. Series A. Mathematical, Physical, and Engineering Sciences, 185(1001), 144–175.
https://doi.org/10.1098/rspa.1946.0010
Steinbrecht, W., Claude, H., Köhler, U., & Hoinka, K. P. (1998).
Correlations between tropopause height and total ozone: Implications for long-term changes. Journal of Geophysical Research:
Atmospheres, 103(D15), 19183–19192. https://doi.org/10.1029/
98JD01929
Hommel, R., Eichmann, K. U., Aschmann, J., Bramstedt, K.,
Weber, M., von Savigny, C., Richter, A., Rozanov, A., Wittrock,
F., Khosrawi, F., Bauer, R., & Burrows, J. P. (2014). Chemical
ozone loss and ozone mini-hole event during the Arctic winter
2010/2011 as observed by SCIAMACHY and GOME-2. Atmospheric Chemistry and Physics, 14(7), 3247–3276. https://doi.org/
10.5194/acp-14-3247-2014
Fountoulakis, I., Diémoz, H., Siani, A. M., di Sarra, A., Meloni,
D., & Sferlazzo, D. M. (2021). Variability and trends in surface
solar spectral ultraviolet irradiance in Italy: On the influence of
geopotential height and lower-stratospheric ozone. Atmospheric
Chemistry and Physics, 21(24), 18689–18705. https://doi.org/
10.5194/acp-21-18689-2021
Lin, P., Paynter, D., Ming, Y., & Ramaswamy, V. (2017).
Changes of the tropical tropopause layer under global warming.
Journal of Climate, 30(4), 1245–1258. https://doi.org/10.1175/
JCLI-D-16-0457.1
Meng, L., Liu, J., Tarasick, D. W., Randel, W. J., Steiner, A.
K., Wilhelmsen, H., Wang, L., & Haimberger, L. (2021). Continuous rise of the tropopause in the Northern Hemisphere over
1980–2020. Science Advances. https://doi.org/10.1126/sciadv.
abi8065
Ball, W. T., Alsing, J., Mortlock, D. J., Staehelin, J., Haigh, J.
D., Peter, T., Tummon, F., Stübi, R., Stenke, A., Anderson, J.,
Bourassa, A., Davis, S. M., Degenstein, D., Frith, S., Froidevaux,
L., Roth, C., Sofieva, V., Wang, R., Wild, J., Yu, P., et al. (2018).
Evidence for a continuous decline in lower stratospheric
ozone offsetting ozone layer recovery. Atmospheric Chemistry and Physics, 18(2), 1379–1394. https:// doi. org/ 10. 5194/
acp-18-1379-2018
13
984
227. McKenzie, R., Liley, B., Kotkamp, M., Geddes, A., Querel,
R., Stierle, S., Lantz, K., Rhodes, S., & Madronich, S. (2022).
Relationship between ozone and biologically relevant UV at
4 NDACC sites. Photochemical & Photobiological Sciences.
https://doi.org/10.1007/s43630-022-00281-5
228. Booth, C. R., Lucas, T. B., Morrow, J. H., Weiler, C. S., & Penhale, P. A. (1994). The United States National Science Foundation’s polar network for monitoring ultraviolet radiation. In
C. S. Weiler & P. A. Penhale (Eds.), Ultraviolet radiation in
Antarctica: Measurements and biological effects, 62 (pp. 17–37).
American Geophysical Union.
229. Bernhard, G., Booth, C. R., & Ehramjian, J. C. (2004). Version
2 data of the National Science Foundation’s ultraviolet radiation monitoring network: South Pole. Journal of Geophysical
Research: Atmospheres, 109(D21), D21207. https://doi.org/10.
1029/2004jd004937
230. Bernhard, G. H., McKenzie, R. L., Lantz, K., & Stierle, S.
(2022). Updated analysis of data from Palmer Station, Antarctica
(64° S), and San Diego, California (32° N), confirms large effect
of the Antarctic ozone hole on UV radiation. Photochemical &
Photobiological Sciences, 21(3), 373–384. https://doi.org/10.
1007/s43630-022-00178-3
231. Cordero, R. R., Damiani, A., Jorquera, J., Sepúlveda, E., Caballero, M., Fernandez, S., Feron, S., Llanillo, P. J., Carrasco, J.,
Laroze, D., & Labbe, F. (2018). Ultraviolet radiation in the Atacama Desert. Antonie van Leeuwenhoek, 111(8), 1301–1313.
https://doi.org/10.1007/s10482-018-1075-z
232. McKenzie, R. L., & Lucas, R. M. (2018). Reassessing impacts
of extended daily exposure to low level solar UV radiation. Scientific Reports, 8(1), 13805. https:// doi. org/ 10. 1038/
s41598-018-32056-3
233. Kirchhoff, V. W. J. H., Sahai, Y., Casiccia, C. A. R. S., Zamorano,
B. F., & Valderrama, V. V. (1997). Observations of the 1995
ozone hole over Punta Arenas. Chile. Journal of Geophysical
Research, 102(D13), 16109–16120. https:// doi. org/ 10. 1029/
97JD00276
234. du Preez, D. J., Ajtić, J. V., Bencherif, H., Bègue, N., Cadet,
J.-M., & Wright, C. Y. (2019). Spring and summer time ozone
and solar ultraviolet radiation variations over Cape Point. South
Africa. Annales Geophysicae, 37(2), 129–141. https://doi.org/10.
5194/angeo-37-129-2019
235. Bernhard, G. H., Fioletov, V. E., Grooß, J.-U., Ialongo, I., Johnsen, B., Lakkala, K., Manney, G. L., Müller, R., & Svendby, T.
(2021). Ozone and ultraviolet radiation. In State of the Climate
in 2020, Bulletin of the American Meteorological Society. https://
doi.org/10.1175/BAMS-D-21-0086.1
236. Bernhard, G. H., Fioletov, V. E., Grooß, J. U., Ialongo, I., Johnsen, B., Lakkala, K., Manney, G. L., Müller, R., & Svendby,
T. (2020). Record-breaking increases in Arctic solar ultraviolet
radiation caused by exceptionally large ozone depletion in 2020.
Geophysical Research Letters. https:// doi. org/ 10. 1029/ 2020g
l090844
237. Bhartia, P. K., & Wellemeyer, C. W. (2002). TOMS-V8 total O3
algorithm. In OMI Algorithm Theoretical Basis Document Volume II, 15–31. NASA Goddard Space Flight Center Tech. Doc.
ATBD-OMI-02
238. Parra, R., Cadena, E., & Flores, C. (2019). Maximum UV index
records (2010–2014) in Quito (Ecuador) and its trend inferred
from remote sensing data (1979–2018). Atmosphere, 10(12), 787.
https://doi.org/10.3390/atmos10120787
239. Rivas, M., Rojas, E., & Madronich, S. (2008). Aumento Del
Índice Solar Ultravioleta Con La Altura. Ingeniare. Revista chilena de ingeniería. https://doi.org/10.4067/S0718-3305200800
0200013
240. Fountoulakis, I., Zerefos, C. S., Bais, A. F., Kapsomenakis, J.,
Koukouli, M.-E., Ohkawara, N., Fioletov, V., De Backer, H.,
13
Photochemical & Photobiological Sciences (2023) 22:937–989
241.
242.
243.
244.
245.
246.
247.
248.
249.
250.
251.
Lakkala, K., Karppinen, T., & Webb, A. R. (2018). Twentyfive years of spectral UV-B measurements over Canada, Europe
and Japan: Trends and effects from changes in ozone, aerosols,
clouds, and surface reflectivity. Comptes Rendus Geoscience,
350(7), 393–402. https://doi.org/10.1016/j.crte.2018.07.011
Zhang, H., Wang, J., Castro García, L., Zeng, J., Dennhardt, C.,
Liu, Y., & Krotkov, N. A. (2019). Surface erythemal UV irradiance in the continental United States derived from ground-based
and OMI observations: Quality assessment, trend analysis and
sampling issues. Atmospheric Chemistry and Physics, 19(4),
2165–2181. https://doi.org/10.5194/acp-19-2165-2019
Chubarova, N. E., Pastukhova, A. S., Zhdanova, E. Y., Volpert,
E. V., Smyshlyaev, S. P., & Galin, V. Y. (2020). Effects of ozone
and clouds on temporal variability of surface UV radiation and
UV resources over Northern Eurasia derived from measurements
and modeling. Atmosphere, 11(1), 59. https://doi.org/10.3390/
atmos11010059
Chubarova, N. E., Pastukhova, A. S., Galin, V. Y., & Smyshlyaev, S. P. (2018). Long-term variability of UV irradiance in the
Moscow region according to measurement and modeling data.
Izvestiya, Atmospheric and Oceanic Physics, 54(2), 139–146.
https://doi.org/10.1134/s0001433818020056
Aun, M., Eerme, K., Ansko, I., & Aun, M. (2019). Daily, seasonal, and annual characteristics of UV radiation and its influencing factors in Tõravere, Estonia, 2004–2016. Theoretical
and Applied Climatology, 138(1–2), 887–897. https://doi.org/
10.1007/s00704-019-02865-1
Fountoulakis, I., Diémoz, H., Siani, A.-M., Laschewski, G.,
Filippa, G., Arola, A., Bais, A. F., De Backer, H., Lakkala, K.,
Webb, A. R., De Bock, V., Karppinen, T., Garane, K., Kapsomenakis, J., Koukouli, M.-E., & Zerefos, C. S. (2020). Solar UV
irradiance in a changing climate: Trends in Europe and the significance of spectral monitoring in Italy. Environments, 7(1), 1.
https://doi.org/10.3390/environments7010001
Hunter, N., Rendell, R. J., Higlett, M. P., & amp, apos, Hagan, J.
B., & Haylock, R. G. E. (2019). Relationship between erythema
effective UV radiant exposure, total ozone, cloud cover and aerosols in southern England. UK. Atmospheric Chemistry and Physics, 19(1), 683–699. https://doi.org/10.5194/acp-19-683-2019
Torres, O., Bhartia, P. K., Jethva, H., & Ahn, C. (2018). Impact
of the ozone monitoring instrument row anomaly on the
long-term record of aerosol products. Atmospheric Measurement Techniques, 11(5), 2701–2715. https:// doi. org/ 10. 5194/
amt-11-2701-2018
Herman, J., Cede, A., Huang, L., Ziemke, J., Torres, O., Krotkov,
N., Kowalewski, M., & Blank, K. (2020). Global distribution
and 14-year changes in erythemal irradiance, UV atmospheric
transmission, and total column ozone for 2005–2018 estimated from OMI and EPIC observations. Atmospheric Chemistry and Physics, 20(14), 8351–8380. https://doi.org/10.5194/
acp-20-8351-2020
Lindfors, A., & Vuilleumier, L. (2005). Erythemal UV at Davos
(Switzerland), 1926–2003, estimated using total ozone, sunshine
duration, and snow depth. Journal of Geophysical Research.
https://doi.org/10.1029/2004JD005231
Krzyścin, J. W., & Sobolewski, P. S. (2018). Trends in erythemal
doses at the Polish Polar Station, Hornsund, Svalbard based on
the homogenized measurements (1996–2016) and reconstructed
data (1983–1995). Atmospheric Chemistry and Physics, 18(1),
1–11. https://doi.org/10.5194/acp-18-1-2018
Posyniak, M., Szkop, A., Pietruczuk, A., Podgórski, J., &
Krzyścin, J. (2016). The long-term (1964–2014) variability
of aerosol optical thickness and its impact on solar irradiance
based on the data taken at Belsk, Poland. Acta Geophysica, 64(5),
1858–1874. https://doi.org/10.1515/acgeo-2016-0026
Photochemical & Photobiological Sciences (2023) 22:937–989
252. Čížková, K., Láska, K., Metelka, L., & Staněk, M. (2018).
Reconstruction and analysis of erythemal UV radiation time
series from Hradec Králové (Czech Republic) over the past 50
years. Atmospheric Chemistry and Physics, 18(3), 1805–1818.
https://doi.org/10.5194/acp-18-1805-2018
253. Volpert, E. V., & Chubarova, N. E. (2021). Long-term changes
in solar radiation in Northern Eurasia during the warm season
according to measurements and reconstruction model. Russian
Meteorology and Hydrology, 46(8), 507–518. https://doi.org/10.
3103/S1068373921080021
254. Malinović-Milićević, S., Radovanović, M. M., Mijatović, Z., &
Petrović, M. D. (2022). Reconstruction and variability of high
daily erythemal ultraviolet doses and relationship with total
ozone, cloud cover, and albedo in Novi Sad (Serbia). International Journal of Climatology. https://doi.org/10.1002/joc.7803
255. WMO. (2011). Scientific Assessment of Ozone Depletion 2010.
Global Ozone Research and Monitoring Project – Report No. 52.
World Meteorological Organization, Geneva, Switzerland.
256. WMO. (2014). Scientific Assessment of Ozone Depletion 2014.
Global Ozone Research and Monitoring Project – Report No. 55.
World Meteorological Organization, Geneva, Switzerland.
257. Bernhard, G., Dahlback, A., Fioletov, V., Heikkilä, A., Johnsen, B., Koskela, T., Lakkala, K., & Svendby, T. (2013). High
levels of ultraviolet radiation observed by ground-based instruments below the 2011 Arctic ozone hole. Atmospheric Chemistry and Physics, 13(21), 10573–10590. https://doi.org/10.5194/
acp-13-10573-2013
258. Kazadzis, S., Founda, D., Psiloglou, B. E., Kambezidis, H.,
Mihalopoulos, N., Sanchez-Lorenzo, A., Meleti, C., Raptis, P.
I., Pierros, F., & Nabat, P. (2018). Long-term series and trends
in surface solar radiation in Athens. Greece. Atmospheric Chemistry and Physics, 18(4), 2395–2411. https://doi.org/10.5194/
acp-18-2395-2018
259. Wild, M., Wacker, S., Yang, S., & Sanchez-Lorenzo, A. (2021).
Evidence for clear-sky dimming and brightening in Central
Europe. Geophysical Research Letters. https://doi.org/10.1029/
2020gl092216
260. Lamy, K., Portafaix, T., Josse, B., Brogniez, C., Godin-Beekmann, S., Bencherif, H., Revell, L., Akiyoshi, H., Bekki, S.,
Hegglin, M. I., Jöckel, P., Kirner, O., Liley, B., Marecal, V., Morgenstern, O., Stenke, A., Zeng, G., Abraham, N. L., Archibald,
A. T., Butchart, N., et al. (2019). Clear-sky ultraviolet radiation modelling using output from the Chemistry Climate Model
Initiative. Atmospheric Chemistry and Physics, 19(15), 10087–
10110. https://doi.org/10.5194/acp-19-10087-2019
261. Eyring, V., Lamarque, J.-F., Hess, P., Arfeuille, F., Bowman,
K., Chipperfiel, M. P., Duncan, B., Fiore, A., Gettelman, A.,
& Giorgetta, M. A. (2013). Overview of IGAC/SPARC Chemistry-Climate Model Initiative (CCMI) community simulations
in support of upcoming ozone and climate assessments. SPARC
Newsletter, 40(1), 48–66.
262. van Vuuren, D. P., Edmonds, J., Kainuma, M., Riahi, K., Thomson, A., Hibbard, K., Hurtt, G. C., Kram, T., Krey, V., Lamarque,
J.-F., Masui, T., Meinshausen, M., Nakicenovic, N., Smith, S.
J., & Rose, S. K. (2011). The representative concentration pathways: An overview. Climatic Change, 109(1–2), 5–31. https://
doi.org/10.1007/s10584-011-0148-z
263. Kinne, S., O’Donnel, D., Stier, P., Kloster, S., Zhang, K.,
Schmidt, H., Rast, S., Giorgetta, M., Eck, T. F., & Stevens, B.
(2013). MAC-v1: A new global aerosol climatology for climate
studies. Journal of Advances in Modeling Earth Systems, 5(4),
704–740. https://doi.org/10.1002/jame.20035
264. Pastukhova, A. S., Chubarova, N. E., Zhdanova, Y. Y., Galin,
V. Y., & Smyshlyaev, S. P. (2019). Numerical simulation of
variations in ozone content, erythemal ultraviolet radiation, and
ultraviolet resources over Northern Eurasia in the 21st century.
985
265.
266.
267.
268.
269.
270.
271.
272.
273.
274.
275.
276.
Izvestiya, Atmospheric and Oceanic Physics, 55(3), 242–250.
https://doi.org/10.1134/s0001433819030058
Eleftheratos, K., Kapsomenakis, J., Zerefos, C. S., Bais, A. F.,
Fountoulakis, I., Dameris, M., Jöckel, P., Haslerud, A. S., GodinBeekmann, S., Steinbrecht, W., Petropavlovskikh, I., Brogniez,
C., Leblanc, T., Liley, J. B., Querel, R., & Swart, D. P. J. (2020).
Possible effects of greenhouse gases to ozone profiles and DNA
active UV-B irradiance at ground level. Atmosphere, 11(3), 228.
https://doi.org/10.3390/atmos11030228
Setlow, R. B. (1974). The wavelengths in sunlight effective in
producing skin cancer: A theoretical analysis. Proceedings of the
National Academy of Sciences of the United States of America,
71(9), 3363–3366. https://doi.org/10.1073/pnas.71.9.3363
Chiodo, G., Polvani, L. M., Marsh, D. R., Stenke, A., Ball, W.,
Rozanov, E., Muthers, S., & Tsigaridis, K. (2018). The response
of the ozone layer to quadrupled CO2 concentrations. Journal of Climate, 31(10), 3893–3907. https:// doi. org/ 10. 1175/
jcli-d-17-0492.1
IPCC. (2014). Climate change 2014: synthesis report. Contribution of working groups I, II and III to the fifth assessment
report of the Intergovernmental Panel on Climate Change, IPCC,
Geneva, Switzerland.
Bardeen, C. G., Kinnison, D. E., Toon, O. B., Mills, M. J., Vitt,
F., Xia, L., Jägermeyr, J., Lovenduski, N. S., Scherrer, K. J. N.,
Clyne, M., & Robock, A. (2021). Extreme Ozone Loss Following
Nuclear War Results in Enhanced Surface Ultraviolet Radiation.
Journal of Geophysical Research. https://doi.org/10.1029/2021J
D035079
Tilmes, S., MacMartin, D. G., Lenaerts, J., van Kampenhout,
L., Muntjewerf, L., Xia, L., Harrison, C. S., Krumhardt, K. M.,
Mills, M. J., & Kravitz, B. (2020). Reaching 1.5 and 2.0 °C
global surface temperature targets using stratospheric aerosol
geoengineering. Earth System Dynamics, 11(3), 579–601. https://
doi.org/10.5194/esd-11-579-2020
Tilmes, S., Richter, J. H., Kravitz, B., MacMartin, D. G., Glanville, A. S., Visioni, D., Kinnison, D. E., & Müller, R. (2021).
Sensitivity of total column ozone to stratospheric sulfur injection
strategies. Geophysical Research Letters. https://doi.org/10.1029/
2021GL094058
Tilmes, S., Visioni, D., Jones, A., Haywood, J., Séférian, R.,
Nabat, P., Boucher, O., Bednarz, E. M., & Niemeier, U. (2022).
Stratospheric ozone response to sulfate aerosol and solar dimming climate interventions based on the G6 Geoengineering
Model Intercomparison Project (GeoMIP) simulations. Atmospheric Chemistry and Physics 22(7), 4557–4579. https://doi.org/
10.5194/acp-22-4557-2022
Robrecht, S., Vogel, B., Tilmes, S., & Müller, R. (2021). Potential of future stratospheric ozone loss in the midlatitudes under
global warming and sulfate geoengineering. Atmospheric Chemistry and Physics, 21(4), 2427–2455. https://doi.org/10.5194/
acp-21-2427-2021
Madronich, S., Tilmes, S., Kravitz, B., MacMartin, D., & Richter,
J. (2018). Response of Surface ultraviolet and visible radiation
to stratospheric SO2 injections. Atmosphere, 9(11), 432. https://
doi.org/10.3390/atmos9110432
Durand, M., Murchie, E. H., Lindfors, A. V., Urban, O., Aphalo,
P. J., & Robson, T. M. (2021). Diffuse solar radiation and canopy photosynthesis in a changing environment. Agricultural and
Forest Meteorology. https://doi.org/10.1016/j.agrformet.2021.
108684
De Mazière, M., Thompson, A. M., Kurylo, M. J., Wild, J. D.,
Bernhard, G., Blumenstock, T., Braathen, G. O., Hannigan, J.
W., Lambert, J.-C., Leblanc, T., McGee, T. J., Nedoluha, G.,
Petropavlovskikh, I., Seckmeyer, G., Simon, P. C., Steinbrecht,
W., & Strahan, S. E. (2018). The network for the detection of
atmospheric composition change (NDACC): history, status and
13
986
277.
278.
279.
280.
281.
282.
283.
284.
285.
286.
287.
288.
289.
Photochemical & Photobiological Sciences (2023) 22:937–989
perspectives. Atmospheric Chemistry and Physics, 18(7), 4935–
4964. https://doi.org/10.5194/acp-18-4935-2018
Hülsen, G., Gröbner, J., Bais, A., Blumthaler, M., Diémoz, H.,
Bolsée, D., Diaz, A., Fountoulakis, I., Naranen, E., Schreder, J.,
Stefania, F., & Guerrero, M. V. (2020). Second solar ultraviolet
radiometer comparison campaign UVC-II. Metrologia. https://
doi.org/10.1088/1681-7575/ab74e5
Svendby, T. M., Johnsen, B., Kylling, A., Dahlback, A., Bernhard, G. H., Hansen, G. H., Petkov, B., & Vitale, V. (2021).
GUV long-term measurements of total ozone column and effective cloud transmittance at three Norwegian sites. Atmospheric
Chemistry and Physics, 21(10), 7881–7899. https://doi.org/10.
5194/acp-21-7881-2021
Foster, T. M., Weide, E. L., Niedzwiedz, A., Duffert, J., & Seckmeyer, G. (2021). Characterization of the angular response of a
multi-directional spectroradiometer for measuring spectral radiance. EPJ Techniques and Instrumentation. https://doi.org/10.
1140/epjti/s40485-021-00069-4
Johnsen, B., Kjeldstad, B., Aalerud, T. N., Nilsen, L. T., Schreder,
J., Blumthaler, M., Bernhard, G., Topaloglou, C., Meinander, O.,
Bagheri, A., Slusser, J. R., & Davis, J. (2008). Intercomparison
and harmonization of UV Index measurements from multiband
filter radiometers. Journal of Geophysical Research. https://doi.
org/10.1029/2007jd009731
Egli, L., Gröbner, J., Hülsen, G., Bachmann, L., Blumthaler, M.,
Dubard, J., Khazova, M., Kift, R., Hoogendijk, K., Serrano, A.,
Smedley, A., & Vilaplana, J. M. (2016). Quality assessment of
solar UV irradiance measured with array spectroradiometers.
Atmospheric Measurement Techniques, 9(4), 1553–1567. https://
doi.org/10.5194/amt-9-1553-2016
Zuber, R., Ribnitzky, M., Tobar, M., Lange, K., Kutscher, D.,
Schrempf, M., Niedzwiedz, A., & Seckmeyer, G. (2018). Global
spectral irradiance array spectroradiometer validation according
to WMO. Measurement Science and Technology. https://doi.org/
10.1088/1361-6501/aada34
González, C., Vilaplana, J. M., Bogeat, J. A., & Serrano, A.
(2022). Comparison of global UV spectral irradiance measurements between a BTS CCD-array and a Brewer spectroradiometer. Atmospheric Measurement Techniques, 15(13), 4125–4133.
https://doi.org/10.5194/amt-15-4125-2022
Allen, M. W., Swift, N., Nield, K. M., Liley, B., & McKenzie,
R. L. (2020). Use of electronic UV dosimeters in measuring personal UV exposures and public health education. Atmosphere.
https://doi.org/10.3390/atmos11070744
Diffey, B. (2020). The early days of personal solar ultraviolet
dosimetry. Atmosphere, 11(2), 125. https:// doi. org/ 10. 3390/
atmos11020125
Downs, N. J., Axelsen, T., Parisi, A. V., Schouten, P. W., & Dexter, B. R. (2020). Measured UV exposures of ironman, sprint and
olympic-distance triathlon competitors. Atmosphere. https://doi.
org/10.3390/atmos11050440
Cai, S., Zuo, C., Zhang, J., Liu, H., & Fang, X. (2021). A paperbased wearable photodetector for simultaneous UV intensity and
dosage measurement. Advanced Functional Materials. https://
doi.org/10.1002/adfm.202100026
Seckmeyer, G., Klingebiel, M., Riechelmann, S., Lohse, I.,
McKenzie, R. L., Ben Liley, J., Allen, M. W., Siani, A.-M., &
Casale, G. R. (2012). A critical assessment of two types of personal UV dosimeters. Photochemistry and Photobiology, 88(1),
215–222. https://doi.org/10.1111/j.1751-1097.2011.01018.x
Gröbner, J., Schreder, J., Kazadzis, S., Bais, A. F., Blumthaler,
M., Görts, P., Tax, R., Koskela, T., Seckmeyer, G., Webb, A. R.,
& Rembges, D. (2005). Traveling reference spectroradiometer for
routine quality assurance of spectral solar ultraviolet irradiance
measurements. Applied Optics, 44(25), 5321–5331. https://doi.
org/10.1364/AO.44.005321
13
290. Cadet, J.-M., Portafaix, T., Bencherif, H., Lamy, K., Brogniez,
C., Auriol, F., Metzger, J.-M., Boudreault, L.-E., & Wright, C.
Y. (2020). Inter-comparison campaign of solar UVR instruments
under clear sky conditions at Reunion Island (21°S, 55°E). International Journal of Environmental Research and Public Health.
https://doi.org/10.3390/ijerph17082867
291. Fountoulakis, I., Diémoz, H., Siani, A. M., Hülsen, G., & Gröbner, J. (2020). Monitoring of solar spectral ultraviolet irradiance
in Aosta. Italy. Earth System Science Data, 12(4), 2787–2810.
https://doi.org/10.5194/essd-12-2787-2020
292. Schinke, C., Pollex, H., Hinken, D., Wolf, M., Bothe, K., Kröger,
I., Nevas, S., & Winter, S. (2020). Calibrating spectrometers for
measurements of the spectral irradiance caused by solar radiation. Metrologia. https://doi.org/10.1088/1681-7575/abafc5
293. Mayer, B., & Kylling, A. (2005). Technical note: The libRadtran
software package for radiative transfer calculations–description
and examples of use. Atmospheric Chemistry and Physics, 5,
1855–1877. https://doi.org/10.5194/acp-5-1855-2005
294. Ricchiazzi, P., Yang, S., Gautier, C., & Sowle, D. (1998).
SBDART: A research and teaching software tool for plane-parallel radiative transfer in the Earth’s atmosphere. Bulletin of the
American Meteorological Society, 79(10), 2101–2114. https://
doi.org/10.1175/1520-0477(1998)079%3c2101:Sarats%3e2.0.
Co;2
295. Gueymard, C. A. (2019). The SMARTS spectral irradiance
model after 25 years: New developments and validation of reference spectra. Solar Energy, 187, 233–253. https://doi.org/10.
1016/j.solener.2019.05.048
296. Madronich, S., & Flocke, S. (1997). Theoretical estimation of
biologically effective UV radiation at the Earth’s surface. In
C. S. Zerefos & A. F. Bais (Eds.), Solar Ultraviolet RadiationModeling, Measurements & Effects, 52 (pp. 23–48). Springer.
297. Madronich, S. (1992). Implications of recent total atmospheric
ozone measurements for biologically active ultraviolet radiation reaching the earth’s surface. Geophysical Research Letters,
19(1), 37–40. https://doi.org/10.1029/91GL02954
298. Emde, C., & Mayer, B. (2007). Simulation of solar radiation
during a total eclipse: A challenge for radiative transfer. Atmospheric Chemistry and Physics, 7(9), 2259–2270. https://doi.org/
10.5194/acp-7-2259-2007
299. Ockenfuß, P., Emde, C., Mayer, B., & Bernhard, G. (2020).
Accurate 3-D radiative transfer simulation of spectral solar irradiance during the total solar eclipse of 21 August 2017. Atmospheric Chemistry and Physics, 20(4), 1961–1976. https://doi.org/
10.5194/acp-20-1961-2020
300. Noebauer, U. M., & Sim, S. A. (2019). Monte Carlo radiative
transfer. Living Reviews in Computational Astrophysics. https://
doi.org/10.1007/s41115-019-0004-9
301. Gueymard, C. A. (2018). Revised composite extraterrestrial spectrum based on recent solar irradiance observations. Solar Energy,
169, 434–440. https://doi.org/10.1016/j.solener.2018.04.067
302. Bak, J., Coddington, O., Liu, X., Chance, K., Lee, H.-J., Jeon,
W., Kim, J.-H., & Kim, C.-H. (2021). Impact of using a new
high-resolution solar reference spectrum on OMI ozone profile
retrievals. Remote Sensing. https://doi.org/10.3390/rs14010037
303. Schmid, B., & Wehrli, C. (1995). Comparison of Sun photometer
calibration by use of the Langley technique and the standard
lamp. Applied Optics, 34(21), 4500–4512. https://doi.org/10.
1364/ao.34.004500
304. Gröbner, J., Kröger, I., Egli, L., Hülsen, G., Riechelmann, S.,
& Sperfeld, P. (2017). The high-resolution extraterrestrial solar
spectrum (QASUMEFTS) determined from ground-based solar
irradiance measurements. Atmospheric Measurement Techniques,
10, 3375–3383. https://doi.org/10.5194/amt-10-3375-2017
305. Richard, E., Harber, D., Coddington, O., Drake, G., Rutkowski,
J., Triplett, M., Pilewskie, P., & Woods, T. (2020). SI-traceable
Photochemical & Photobiological Sciences (2023) 22:937–989
306.
307.
308.
309.
310.
311.
312.
313.
314.
315.
316.
317.
spectral irradiance radiometric characterization and absolute calibration of the TSIS-1 Spectral Irradiance Monitor (SIM). Remote
Sensing. https://doi.org/10.3390/rs12111818
Coddington, O. M., Richard, E. C., Harber, D., Pilewskie, P.,
Woods, T. N., Chance, K., Liu, X., & Sun, K. (2021). The TSIS-1
Hybrid Solar Reference Spectrum. Geophysical Research Letters.
https://doi.org/10.1029/2020gl091709
Arola, A., Wandji Nyamsi, W., Lipponen, A., Kazadzis, S., Krotkov, N. A., & Tamminen, J. (2021). Rethinking the correction
for absorbing aerosols in the OMI- and TROPOMI-like surface
UV algorithms. Atmospheric Measurement Techniques, 14(7),
4947–4957. https://doi.org/10.5194/amt-14-4947-2021
Heath, D. F., Mateer, C. L., & Krueger, A. J. (1973). The Nimbus-4 Backscatter Ultraviolet (BUV) atmospheric ozone experiment — tow years’ operation. Pure and Applied Geophysics,
106–108(1), 1238–1253. https://doi.org/10.1007/BF00881076
Frith, S. M., Kramarova, N. A., Stolarski, R. S., McPeters, R.
D., Bhartia, P. K., & Labow, G. J. (2014). Recent changes in
total column ozone based on the SBUV Version 8.6 Merged
Ozone Data Set. Journal of Geophysical Research: Atmospheres,
119(16), 9735–9751. https://doi.org/10.1002/2014JD021889
Krotkov, N. A., Bhartia, P. K., Herman, J. R., Fioletov, V., &
Kerr, J. (1998). Satellite estimation of spectral surface UV irradiance in the presence of tropospheric aerosols: 1. Cloud-free case.
Journal of Geophysical Research, 103(D8), 8779–8793. https://
doi.org/10.1029/98jd00233
Krotkov, N. A., Herman, J. R., Bhartia, P. K., Fioletov, V., &
Ahmad, Z. (2001). Satellite estimation of spectral surface UV
irradiance: 2. Effects of homogeneous clouds and snow. Journal of Geophysical Research: Atmospheres, 106(D11), 11743–
11759. https://doi.org/10.1029/2000jd900721
Burrows, J. P., Weber, M., Buchwitz, M., Rozanov, V., LadstätterWeißenmayer, A., Richter, A., DeBeek, R., Hoogen, R., Bramstedt, K., Eichmann, K.-U., Eisinger, M., & Perner, D. (1999). The
Global Ozone Monitoring Experiment (GOME): Mission concept and first scientific results. Journal of the Atmospheric Sciences, 56(2), 151–175. https://doi.org/10.1175/1520-0469(1999)
056%3c0151:Tgomeg%3e2.0.Co;2
Kujanpää, J., & Kalakoski, N. (2015). Operational surface UV
radiation product from GOME-2 and AVHRR/3 data. Atmospheric Measurement Techniques, 8(10), 4399–4414. https://doi.
org/10.5194/amt-8-4399-2015
Levelt, P. F., Joiner, J., Tamminen, J., Veefkind, J. P., Bhartia,
P. K., Stein Zweers, D. C., Duncan, B. N., Streets, D. G., Eskes,
H., van der A, R., McLinden, C., Fioletov, V., Carn, S., de Laat,
J., DeLand, M., Marchenko, S., McPeters, R., Ziemke, J., Fu, D.,
Liu, X., et al. (2018). The Ozone Monitoring Instrument: overview of 14 years in space. Atmospheric Chemistry and Physics,
18(8), 5699–5745. https://doi.org/10.5194/acp-18-5699-2018
Herman, J., Huang, L., McPeters, R., Ziemke, J., Cede, A., &
Blank, K. (2018). Synoptic ozone, cloud reflectivity, and erythemal irradiance from sunrise to sunset for the whole earth as
viewed by the DSCOVR spacecraft from the earth–sun Lagrange
1 orbit. Atmospheric Measurement Techniques, 11(1), 177–194.
https://doi.org/10.5194/amt-11-177-2018
Brogniez, C., Auriol, F., Deroo, C., Arola, A., Kujanpää, J.,
Sauvage, B., Kalakoski, N., Pitkänen, M. R. A., Catalfamo, M.,
Metzger, J. M., Tournois, G., & Da Conceicao, P. (2016). Validation of satellite-based noontime UVI with NDACC ground-based
instruments: Influence of topography, environment and satellite
overpass time. Atmospheric Chemistry and Physics, 16(23),
15049–15074. https://doi.org/10.5194/acp-16-15049-2016
Arola, A., Kazadzis, S., Lindfors, A., Krotkov, N., Kujanpää, J.,
Tamminen, J., Bais, A., di Sarra, A., Villaplana, J. M., Brogniez,
C., Siani, A. M., Janouch, M., Weihs, P., Webb, A., Koskela, T.,
Kouremeti, N., Meloni, D., Buchard, V., Auriol, F., Ialongo, I.,
987
318.
319.
320.
321.
322.
323.
324.
325.
326.
327.
et al. (2009). A new approach to correct for absorbing aerosols in
OMI UV. Geophysical Research Letters, 36(22), L22805. https://
doi.org/10.1029/2009gl041137
Bernhard, G., Arola, A., Dahlback, A., Fioletov, V., Heikkilä, A.,
Johnsen, B., Koskela, T., Lakkala, K., Svendby, T., & Tamminen,
J. (2015). Comparison of OMI UV observations with groundbased measurements at high northern latitudes. Atmospheric
Chemistry and Physics, 15(13), 7391–7412. https://doi.org/10.
5194/acp-15-7391-2015
Kosmopoulos, P. G., Kazadzis, S., Schmalwieser, A. W., Raptis, P. I., Papachristopoulou, K., Fountoulakis, I., Masoom, A.,
Bais, A. F., Bilbao, J., Blumthaler, M., Kreuter, A., Siani, A.
M., Eleftheratos, K., Topaloglou, C., Gröbner, J., Johnsen, B.,
Svendby, T. M., Vilaplana, J. M., Doppler, L., Webb, A. R., et al.
(2021). Real-time UV index retrieval in Europe using Earth
observation-based techniques: System description and quality assessment. Atmospheric Measurement Techniques, 14(8),
5657–5699. https://doi.org/10.5194/amt-14-5657-2021
Vuilleumier, L., Harris, T., Nenes, A., Backes, C., & Vernez, D.
(2021). Developing a UV climatology for public health purposes
using satellite data. Environment International. https://doi.org/
10.1016/j.envint.2020.106177
Fisher, D. (2020). Mission Status for Earth Science Constellation.
In NASA Goddard Space Flight Center Tech. Doc. Document ID
20205007514, 1–41: Goddard Space Flight Center, Greenbelt,
Maryland.
Flynn, L., Long, C., Wu, X., Evans, R., Beck, C. T., Petropavlovskikh, I., McConville, G., Yu, W., Zhang, Z., & Niu, J. (2014).
Performance of the ozone mapping and profiler suite (OMPS)
products. Journal of Geophysical Research: Atmospheres,
119(10), 6181–6195. https://doi.org/10.1002/2013JD020484
Lindfors, A. V., Kujanpää, J., Kalakoski, N., Heikkilä, A., Lakkala, K., Mielonen, T., Sneep, M., Krotkov, N. A., Arola, A., &
Tamminen, J. (2018). The TROPOMI surface UV algorithm.
Atmospheric Measurement Techniques, 11(2), 997–1008. https://
doi.org/10.5194/amt-11-997-2018
Cortesi, U., Ceccherini, S., Del Bianco, S., Gai, M., Tirelli, C.,
Zoppetti, N., Barbara, F., Bonazountas, M., Argyridis, A., Bós,
A., Loenen, E., Arola, A., Kujanpää, J., Lipponen, A., Wandji
Nyamsi, W., van der A, R., van Peet, J., Tuinder, O., Farruggia,
V., Masini, A., et al. (2018). Advanced Ultraviolet Radiation and
Ozone Retrieval for Applications (AURORA): A project overview. Atmosphere, 9(11), 454. https:// doi. org/ 10. 3390/ atmos
9110454
Lakkala, K., Kujanpää, J., Brogniez, C., Henriot, N., Arola, A.,
Aun, M., Auriol, F., Bais, A. F., Bernhard, G., De Bock, V.,
Catalfamo, M., Deroo, C., Diémoz, H., Egli, L., Forestier, J.-B.,
Fountoulakis, I., Garcia, R. D., Gröbner, J., Hassinen, S., Heikkilä, A., et al. (2020). Validation of the TROPOspheric Monitoring Instrument (TROPOMI) surface UV radiation product.
Atmospheric Measurement Techniques 13(12), 6999–7024.
https://doi.org/10.5194/amt-13-6999-2020
Tanskanen, A., Lindfors, A., Maatta, A., Krotkov, N., Herman, J.,
Kaurola, J., Koskela, T., Lakkala, K., Fioletov, V., Bernhard, G.,
McKenzie, R., Kondo, Y., O’Neill, M., Slaper, H., den Outer, P.,
Bais, A. F., & Tamminen, J. (2007). Validation of daily erythemal doses from ozone monitoring instrument with ground-based
UV measurement data. Journal of Geophysical Research. https://
doi.org/10.1029/2007JD008830
Zempila, M. M., Fountoulakis, I., Taylor, M., Kazadzis, S., Arola,
A., Koukouli, M. E., Bais, A., Meleti, C., & Balis, D. (2018).
Validation of OMI erythemal doses with multi-sensor groundbased measurements in Thessaloniki, Greece. Atmospheric Environment, 183, 106–121. https://doi.org/10.1016/j.atmosenv.2018.
04.012
13
988
328. Lipponen, A., Ceccherini, S., Cortesi, U., Gai, M., Keppens,
A., Masini, A., Simeone, E., Tirelli, C., & Arola, A. (2020).
Advanced ultraviolet radiation and ozone retrieval for applications—surface ultraviolet radiation products. Atmosphere. https://
doi.org/10.3390/atmos11040324
329. Zhao, R., & He, T. (2022). Estimation of 1-km tesolution all-sky
instantaneous erythemal UV-B with MODIS data based on a
deep learning method. Remote Sensing. https://doi.org/10.3390/
rs14020384
330. Krzyścin, J. W., Guzikowski, J., Pietruczuk, A., & Sobolewski,
P. S. (2019). Improvement of the 24 hr forecast of surface UV
radiation using an ensemble approach. Meteorological Applications. https://doi.org/10.1002/met.1865
331. Ahmed, A. A. M., Ahmed, M. H., Saha, S. K., Ahmed, O.,
& Sutradhar, A. (2022). Optimization algorithms as training
approach with hybrid deep learning methods to develop an
ultraviolet index forecasting model. Stochastic Environmental
Research and Risk Assessment, 36(10), 3011–3039. https://doi.
org/10.1007/s00477-022-02177-3
332. Scragg, R. K. R., Stewart, A. W., McKenzie, R. L., Reeder, A. I.,
Liley, J. B., & Allen, M. W. (2016). Sun exposure and 25-hydroxyvitamin D3 levels in a community sample: Quantifying the association with electronic dosimeters. Journal of Exposure Science
& Environmental Epidemiology, 27(5), 471–477. https://doi.org/
10.1038/jes.2016.51
333. Godar, D. E. (2005). UV Doses Worldwide. Photochemistry and
Photobiology. https://doi.org/10.1111/j.1751-1097.2005.tb014
38.x
334. Schmalwieser, A. W., & Siani, A. M. (2018). Review on nonoccupational personal solar UV exposure measurements. Photochemistry and Photobiology, 94(5), 900–915. https://doi.org/10.
1111/php.12946
335. Schmalwieser, A. W., Casale, G. R., Colosimo, A., Schmalwieser, S. S., & Siani, A. M. (2021). Review on occupational
personal solar UV exposure measurements. Atmosphere. https://
doi.org/10.3390/atmos12020142
336. Backes, C., Religi, A., Moccozet, L., Vuilleumier, L., Vernez,
D., & Bulliard, J. L. (2018). Facial exposure to ultraviolet radiation: Predicted sun protection effectiveness of various hat styles.
Photodermatology, Photoimmunology and Photomedicine, 34(5),
330–337. https://doi.org/10.1111/phpp.12388
337. Backes, C., Religi, A., Moccozet, L., Behar-Cohen, F., Vuilleumier, L., Bulliard, J.-L., & Vernez, D. (2019). Sun exposure to the eyes: Predicted UV protection effectiveness of
various sunglasses. Journal of Exposure Science & Environmental Epidemiology, 29(6), 753–764. https://doi.org/10.1038/
s41370-018-0087-0
338. Mims, F. M., III., McGonigle, A. J. S., Wilkes, T. C., Parisi, A.
V., Grant, W. B., Cook, J. M., & Pering, T. D. (2019). Measuring
and visualizing solar UV for a wide range of atmospheric conditions on Hawai’i island. International Journal of Environmental
Research and Public Health, 16(6), 997. https://doi.org/10.3390/
ijerph16060997
339. Wester, U., Pagels, P., & Boldemann, C. (2018). Solar ultraviolet
radiation exposure proxy-estimated by sky view fish-eye photography—potentials and limitations from an exploratory correlation
study. Photochemistry and Photobiology, 95(2), 656–661. https://
doi.org/10.1111/php.13027
340. Schmalwieser, A. W., Lohr, M. A., Daly, S. M., & Williams, J. D.
(2022). Modeling acute and cumulative erythemal sun exposure
on vulnerable body sites during beach vacations utilizing behavior-encoded 3D body models. Photochemical & Photobiological
Sciences. https://doi.org/10.1007/s43630-022-00293-1
341. Henning, A., Downs, J., & N., & Vanos, J. K. (2021). Wearable ultraviolet radiation sensors for research and personal use.
13
Photochemical & Photobiological Sciences (2023) 22:937–989
342.
343.
344.
345.
346.
347.
348.
349.
350.
351.
352.
353.
354.
355.
International Journal of Biometeorology, 66(3), 627–640. https://
doi.org/10.1007/s00484-021-02216-8
Huang, X., & Chalmers, A. N. (2021). Review of wearable and
portable sensors for monitoring personal solar UV exposure.
Annals of Biomedical Engineering, 49(3), 964–978. https://doi.
org/10.1007/s10439-020-02710-x
Zou, W., Sastry, M., Gooding, J. J., Ramanathan, R., & Bansal, V.
(2020). Recent advances and a roadmap to wearable UV sensor
technologies. Advanced Materials Technologies. https://doi.org/
10.1002/admt.201901036
Wilson, D. J., Martín-Martínez, F. J., & Deravi, L. F. (2022).
Wearable light sensors based on unique features of a natural biochrome. ACS Sensors, 7(2), 523–533. https://doi.org/10.1021/
acssensors.1c02342
Wang, J., Jeevarathinam, A. S., Jhunjhunwala, A., Ren, H.,
Lemaster, J., Luo, Y., Fenning, D. P., Fullerton, E. E., & Jokerst,
J. V. (2018). A wearable colorimetric dosimeter to monitor sunlight exposure. Advanced Materials Technologies. https://doi.org/
10.1002/admt.201800037
Terenetskaya, I. (2019). How to provide adequate UV dosimetry
to avoid Vitamin D deficiency and related neuropsychiatric disorders. Archives in Neurology & Neuroscience. https://doi.org/
10.33552/ANN.2019.06.000629
de Troya Martín, M., Sánchez, N. B., Harana, C. G., Alarcón
Leiva, M. C., Arjona, J. A., Ruiz, F. R., & de Gálvez Aranda.
(2021). Beach lifeguards′ sun exposure and sun protection in
Spain. Safety and Health at Work, 12(2), 244–248. https://doi.
org/10.1016/j.shaw.2020.10.003
Robinson, J. K., Durst, D. A., Gray, E., Kwasny, M., Heo, S. Y.,
Banks, A., & Rogers, J. A. (2020). Sun exposure reduction by
melanoma survivors with wearable sensor providing real-time
UV exposure and daily text messages with structured goal setting. Archives of Dermatological Research, 313(8), 685–694.
https://doi.org/10.1007/s00403-020-02163-1
Modenese, A., Korpinen, L., & Gobba, F. (2018). Solar radiation
exposure and outdoor work: an underestimated occupational risk.
International Journal of Environmental Research and Public
Health. https://doi.org/10.3390/ijerph15102063
Robinson, J. K., Patel, S., Heo, S. Y., Gray, E., Lim, J., Kwon,
K., Christiansen, Z., Model, J., Trueb, J., Banks, A., Kwasny,
M., & Rogers, J. A. (2021). Real-time UV measurement with a
sun protection system for warning young adults about sunburn:
Prospective cohort study. JMIR mHealth and uHealth. https://
doi.org/10.2196/25895
Moldovan, H. R., Wittlich, M., John, S. M., Brans, R., Tiplica,
G. S., Salavastru, C., Voidazan, S. T., Duca, R. C., Fugulyan, E.,
Horvath, G., Alexa, A., & Butacu, A. I. (2020). Exposure to solar
UV radiation in outdoor construction workers using personal
dosimetry. Environmental Research. https://doi.org/10.1016/j.
envres.2019.108967
Thieden, E., Holm-Schou, A.-S.S., Philipsen, P. A., Heydenreich,
J., & Wulf, H. C. (2019). Adult UVR exposure changes with life
stage – a 14-year follow-up study using personal electronic UVR
dosimeters. Photochemical & Photobiological Sciences, 18(2),
467–476. https://doi.org/10.1039/c8pp00365c
De Castro-Maqueda, G., Gutierrez-Manzanedo, J. V., PonceGonzález, J. G., Fernandez-Santos, J. R., Linares-Barrios, M.,
& De Troya-Martín, M. (2019). Sun protection habits and sunburn in elite aquatics athletes: Surfers, windsurfers and olympic
sailors. Journal of Cancer Education, 35(2), 312–320. https://
doi.org/10.1007/s13187-018-1466-x
Nakagawara, V. B., Montgomery, R. W., & Marshall, W. J.
(2007). Optical radiation transmittance of aircraft windscreeens
and pilot vision. Office of Aerospace Medicine.
International Commission on Non-Ionizing Radiation Protection.
(2004). Guidelines on limits of exposure to ultraviolet radiation
Photochemical & Photobiological Sciences (2023) 22:937–989
356.
357.
358.
359.
360.
361.
362.
363.
of wavelengths between 180 nm and 400 nm (incoherent optical
radiation). Health Physics, 87(2), 171–186.
Baczynska, K. A., Brown, S., Chorley, A. C., Lyachev, A.,
Wittlich, M., & Khazova, M. (2020). Measurements of UV-A
exposure of commercial pilots using Genesis-UV dosimeters.
Atmosphere. https://doi.org/10.3390/atmos11050475
Turner, J., Igoe, D., Parisi, A. V., McGonigle, A. J., Amar, A., &
Wainwright, L. (2020). A review on the ability of smartphones
to detect ultraviolet (UV) radiation and their potential to be used
in UV research and for public education purposes. Science of
the Total Environment. https://doi.org/10.1016/j.scitotenv.2019.
135873
Salvadori, G., Leccese, F., Lista, D., Burattini, C., & Bisegna, F.
(2020). Use of smartphone apps to monitor human exposure to
solar radiation: Comparison between predicted and measured UV
index values. Environmental Research. https://doi.org/10.1016/j.
envres.2020.109274
Kwon, K., Heo, S. Y., Yoo, I., Banks, A., Chan, M., Lee, J. Y.,
Park, J. B., Kim, J., & Rogers, J. A. (2019). Miniaturized, lightadaptive, wireless dosimeters autonomously monitor exposure to
electromagnetic radiation. Science Advances. https://doi.org/10.
1126/sciadv.aay2462
Harris, T. C., Vuilleumier, L., Backes, C., Nenes, A., & Vernez,
D. (2021). Satellite-based personal UV dose estimation. Atmosphere. https://doi.org/10.3390/atmos12020268
Krzyścin, J. W., Lesiak, A., Narbutt, J., Sobolewski, P., &
Guzikowski, J. (2018). Perspectives of UV nowcasting to monitor
personal pro-health outdoor activities. Journal of Photochemistry and Photobiology B: Biology, 184, 27–33. https://doi.org/10.
1016/j.jphotobiol.2018.05.012
Kohli, I., Zubair, R., Lyons, A. B., Nahhas, A. F., Braunberger,
T. L., Mokhtari, M., Ruvolo, E., Lim, H. W., & Hamzavi, I. H.
(2019). Impact of long-wavelength ultraviolet al and visible light
on light-skinned individuals. Photochemistry and Photobiology,
95(6), 1285–1287. https://doi.org/10.1111/php.13143
Diffey, B., & Osterwalder, U. (2017). Labelled sunscreen SPFs
may overestimate protection in natural sunlight. Photochemical
989
364.
365.
366.
367.
368.
369.
370.
371.
& Photobiological Sciences, 16(10), 1519–1523. https://doi.org/
10.1039/C7PP00260B
MacLaughlin, J. A., Anderson, R. R., & Holick, M. F. (1982).
Spectral character of sunlight modulates photosynthesis of
previtamin D3 and its photoisomers in human skin. Science,
216(4549), 1001–1003. https://doi.org/10.1126/science.6281884
Bouillon, R., Eisman, J., Garabedian, M., Holick, M., Kleinschmidt, J., Suda, T., Terenetskaya, I., & Webb, A. (2006). Action
spectrum for the production of previtamin D3 in human skin.
Commission Internationale de l’Eclairage, Austria.
Norval, M., Björn, L. O., & de Gruijl, F. R. (2010). Is the action
spectrum for the UV-induced production of previtamin D3 in
human skin correct? Photochemical and Photobiological Sciences, 9(1), 11–17. https://doi.org/10.1039/b9pp00012g
Young, A. R., Morgan, K. A., Harrison, G. I., Lawrence, K. P.,
Petersen, B., Wulf, H. C., & Philipsen, P. A. (2021). A revised
action spectrum for vitamin D synthesis by suberythemal
UV radiation exposure in humans in vivo. Proceedings of the
National Academy of Sciences. https:// doi. org/ 10. 1073/ pnas.
2015867118
Bolsée, D., Webb, A. R., Gillotay, D., Dörschel, B., Knuschke,
P., Krins, A., & Terenetskaya, I. (2000). Laboratory facilities and
recommendations for the characterization of biological ultraviolet dosimeters. Applied Optics, 39(16), 2813–2822. https://doi.
org/10.1364/ao.39.002813
Olds, W. (2010). Elucidating the links between UV radiation
and vitamin D synthesis: Using an in vitro model. PhD Thesis,
Queensland University of Technology Brisbane, Australia
van Dijk, A., den Outer, P., van Kranen, H., & Slaper, H. (2016).
The action spectrum for vitamin D3: Initial skin reaction and
prolonged exposure. Photochemical & Photobiological Sciences,
15(7), 896–909. https://doi.org/10.1039/C6PP00034G
Webb, A. R., Terenetskaya, I. P., Holick, M. F., van Dijk, A.,
McKenzie, R. L., Lucas, R. M., Young, A. R., Philipsen, P. A., &
de Gruijl, F. R. (2022). Previtamin D action spectrum: challenging CIE towards a standard. Lighting Research & Technology.
https://doi.org/10.1177/14771535221122937
13