ArticlePDF Available

Abstract and Figures

Observations of excited hydroxyl (OH*) emissions are broadly used for inferring information about atmospheric dynamics and composition. We present several analytical approximations for characterizing the excited hydroxyl layer in the Martian atmosphere. They include the OH* number density at the maximum and the height of the peak, along with the relations for assessing different impacts on the OH* layer under night-time conditions. These characteristics are determined by the ambient temperature, atomic oxygen concentration, and their vertical gradients. The derived relations can be used for the analysis of airglow measurements and the interpretation of their variations.
Content may be subject to copyright.
Remote Sens. 2022, 14, 3866. https://doi.org/10.3390/rs14163866 www.mdpi.com/journal/remotesensing
Article
Simplified Relations for the Martian Night-Time OH* Suitable
for the Interpretation of Observations
Mykhaylo Grygalashvyly 1,*, Dmitry S. Shaposhnikov 2, Alexander S. Medvedev 1, Gerd Reinhold Sonnemann 1
and Paul Hartogh 1
1 Max Planck Institute for Solar System Research, 37077 Goettingen, Germany
2 Moscow Institute of Physics and Technology, 141701 Moscow, Russia
* Correspondence: gryga@iap-kborn.de
Abstract: Observations of excited hydroxyl (OH*) emissions are broadly used for inferring infor-
mation about atmospheric dynamics and composition. We present several analytical approxima-
tions for characterizing the excited hydroxyl layer in the Martian atmosphere. They include the OH*
number density at the maximum and the height of the peak, along with the relations for assessing
different impacts on the OH* layer under night-time conditions. These characteristics are deter-
mined by the ambient temperature, atomic oxygen concentration, and their vertical gradients. The
derived relations can be used for the analysis of airglow measurements and the interpretation of
their variations.
Keywords: Mars; excited hydroxyl; Martian atmosphere; airglow; OH*; nightglow
1. Introduction
Hydroxyl molecules in excited states (OH*) produce airglow in visible and near-IR
bands. Excited hydroxyl in the vibrationally excited state originates from the reaction of
ozone with atomic hydrogen; then, it can be deactivated by collisions with other mole-
cules and atoms, chemically removed by reaction with atomic oxygen, or emit a photon
by spontaneous emission. The distribution and abundances of hydroxyl are very sensitive
to atmospheric dynamics, thermodynamics, and photochemistry. Therefore, airglow
measurements provide a useful tool for studying these processes. In the terrestrial atmos-
phere, observations of emissions of OH* are broadly used to obtain information about
tides [1,2], planetary waves [3,4], gravity waves [5–7], and quasi-biennial oscillation [8].
These emissions are also utilized for studying sudden stratospheric warming events
[9,10]. Observations of OH* emissions have been used for retrieving temperature trends
and variations induced by the solar cycle, e.g., [1115], and chemical composition in the
mesopause region [1618].
Recently, hydroxyl emissions were found in the atmosphere of Venus [1923] and on
Mars [24]. Future observations open the possibility for similar applications of the emis-
sions at these planets (for example, investigations of waves and tides by airglow observa-
tions and measurements of atomic oxygen concentrations). Commonly, complex photo-
chemical and general circulation models (i.e., non-linear global with interactively coupled
dynamics, chemistry, and radiation) are required for reproducing the behavior of the OH*
layer, the main characteristics of which are the altitude, emission intensity, and the shape.
When interpretating measurements, it is desirable to establish straightforward relations
between these quantities and the ambient temperature, air density, and concentration of
minor species involved in photochemical reactions and induced emissions. Since full so-
lutions are complex, it is not easy to assess the impacts of individual processes and inter-
pret the variabilities. Since the conditions differ between planets, we focus on Mars in this
paper.
Citation:
Grygalashvyly, M.;
Shaposhnikov, D.S.; Medvedev, A.S.;
Sonnemann, G.R. Simplified
Relations for the Martian
Night
-Time OH* Suitable for the
Interpretation of Observations.
Remote Sens.
2022, 14, 3866.
https://doi.org/10.3390/rs14163866
Academ
ic Editors: Lin Li, Yuanzhi
Zhang and Shengbo Chen
Received:
1 July 2022
Accepted:
7 August 2022
Published:
9 August 2022
Publisher’s Note:
MDPI stays neu-
tral with regard to
jurisdictional
claims in published maps and institu-
tional affiliations.
Copyright:
© 2022 by the authors. Li-
censee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and con-
ditions of the Creative Commons At-
trib
ution (CC BY) license (https://cre-
ativecommons.org/licenses/by/4.0/).
Remote Sens. 2022, 14, 3866 2 of 12
Satellite airglow measurements are not sufficiently precise and result in typical errors
in the determination of the layer altitude ~2–3 km. Ground-based observations are re-
stricted to local points and integrated volume emission, which leads to even larger errors
in the determination of the altitude (on Earth, the OH* layer is commonly assumed at 87–
88 km). In order to study the morphology and variability of the layer, we select the con-
centration of OH* at the peak, which is directly proportional to the volume emission, and
the altitude of the maximum as the characteristics of interest. In the next section, we ana-
lytically derive several approximations for these parameters as well as for relative varia-
tions of the OH* layer. In Section 3, we present applications of the derived formulae based
on the input from the Mars Climate Database (MCD) and determine their validity. Con-
clusions are given in Section 4.
2. Analytical Formulae Derivation
The list of photochemical reactions pertinent to hydroxyl in the Martian atmosphere,
along with the corresponding rates, is given in Table 1.
Table 1. List of reactions, nomenclature of reaction rates, quenching coefficients, and spontaneous
emission coefficients used in the paper.
Reactions
Coefficients
References
R1
. H + O
󰇒
OH,…,+ O = 1.4 10470
,…,= 0.47,0.34,0.15,0.03,0.01
[25,26]
R2
. O + O
+ CO
O
+ CO
= 6.1 10
(
298
).
[25]
R3
. O + O2O = 8 10
2060
[25]
R4
. O + OH,..,O+ H (= 9, ,1)= (5.42,4.8,
4.42, 4, 3.77,4.43,3.74,3,3.15)10

[27]
R5
.OH+ CO, O, N, O
OH󰆒+ CO, O, N, O 󰆒󰆒,󰆒,󰆒
See text
[2629]
R6
.OH
OH
󰆒
+ h
󰆒
[30]
The table includes the source reaction for vibrationally excited hydroxyl (R1), the re-
action of chemical removal (R4), the reactions for collisional deactivation (R5), and spon-
taneous emission (R6). Reactions R2 and R3 are related to the ozone balance equation,
which will be used below. This list omits the reaction of hydroperoxy radicals (HO2) with
atomic oxygen because they represent a negligible (or even non-existing) source for the
population of vibrationally excited hydroxyl [3034]. Thus, the starting point of our con-
sideration is the almost complete set of equations for OH*.
Next, we assume that the excited hydroxyl is in a photochemical equilibrium at night
[31]; hence, we can write its concentration as a ratio of production to the loss term. This
allows us to explicitly express the concentration of excited hydroxyl at all excitation levels
(v = 1, …, 9) in the form
[][][]+󰆓[󰆓]
󰆓 []+󰆓[󰆓]
󰆓 []+
+󰆓[󰆓]
󰆓 []+󰆓[󰆓]
󰆓 []+
+󰆓[󰆓]
󰆓
󰇭󰆓󰆓[]

󰆓󰆓 +󰆓󰆓[]

󰆓󰆓 +󰆓󰆓[]

󰆓󰆓 +
+󰆓󰆓[]

󰆓󰆓 +󰆓󰆓

󰆓󰆓 +
+()[]󰇮,󰇡<󰆒
󰆒󰆒<󰇢 (1
)
Remote Sens. 2022, 14, 3866 3 of 12
where v is the vibrational number; fv are the nascent distributions; r1 and r4 are the reaction
rates; and Avv, Bvv, Gvv, and Dvv are the quenching coefficients by carbon dioxide, molec-
ular oxygen, molecular nitrogen, and atomic oxygen, respectively. Hereafter, the square
brackets denote the number density of a particular chemical constituent. Relation (1) can
be simplified by only considering the main processes of production and relaxation,
namely, the reaction of ozone with atomic hydrogen, quenching by carbon dioxide, mo-
lecular oxygen, and molecular nitrogen:
[]󰇧[][]+󰆓[󰆓]
󰆓 []+
+󰆓[󰆓]
󰆓 []+󰆓[󰆓]
󰆓 []󰇨
󰆓󰆓[]

󰆓󰆓 +󰆓󰆓[]

󰆓󰆓 +󰆓󰆓[]

󰆓󰆓 ,󰇡<󰆒
󰆒󰆒<󰇢. (2
)
In (2), we neglected a spontaneous emission and quenching by atomic oxygen be-
cause these processes are weak on Mars. For example, the total spontaneous emission co-
efficients for vibrational levels OHv = 9 and OHv = 1 are E9 = 199.2495 s−1 and E1 = 17.62 s−1,
respectively [30]. On the other hand, [CO2] 1015 cm−3 at 50 km, e.g., [35,36], the collisional
removal rates A9 = 9.1 × 10−11 cm3 s−1, and A1 = 2.9 × 10−13 cm3s−1 [29,31,3739] yield the first
term in the denominator in (1) exceeding 9 × 104 s−1 and 2.9 × 102 s−1 for the corresponding
vibrational numbers. Atomic oxygen concentrations at 50–60 km are around 1091011 cm−3,
e.g., [35,40,41]. Ref. [27] derived for reactive (O + OHv O2 + H) and non-reactive (O +
OHv OHv<v + O) quenching rates by atomic oxygen (at T = 160 K) the values 7.7 × 10−11
cm3 s−1 and 6 × 10−11 cm3 s−1 for v = 9 and 1, respectively. Hence, the corresponding colli-
sional removal rate due to atomic oxygen is less than 8–6 s1 for all the vibrational numbers
and can be neglected.
Following the work of [31], we assume that ozone is in a photochemical equilibrium
in the vicinity of the night-time OH* layer. Then, the balance equation for ozone can be
represented as [][][]=[][]+[][]. (3
)
The share of the reaction of ozone with atomic oxygen in total ozone loss is small
since, for typical temperatures at 50–60 km (~150 K), the reaction rate r3 (~8.7 × 10−18 cm3
s−1) is about 106 times smaller than r1 (~6.1 × 10−12 cm3 s−1), but the atomic hydrogen number
density is smaller than that of atomic oxygen by no more than ~102103 times in this region
[31,35,36,40]. Therefore, the second term on the right-hand side of (3) can be neglected:
[][][][][]. (4
)
The substitution of (4) into the first term in the numerator in (2) gives
[]󰇧[][][]+󰆓[󰆓]
󰆓 []+
+󰆓[󰆓]
󰆓 []+󰆓[󰆓]
󰆓 []󰇨
󰆓󰆓[]

󰆓󰆓 +󰆓󰆓[]

󰆓󰆓 +󰆓󰆓[]

󰆓󰆓 ,󰇡<󰆒
󰆒󰆒<󰇢. (5
)
Molecular oxygen and molecular nitrogen number densities are linearly proportional
to the concentration of carbon dioxide []=[]=,[]=[], where M is the
air number density, and α, β, and χ are the proportionality coefficients at the heights of
the OH* layer, e.g., [35,41]. In the current work, one will find such behavior below in Fig-
ure 1a. This allows us to exclude the dependencies on concentrations [O2] and [N2] and re-
arrange (5):
[][]+[󰆓]󰆒
󰆓
󰆒󰆒

󰆓󰆓 ,󰇡<󰆒
󰆒󰆒<󰇢, (6
)
where 󰆓
󰆓 =󰆓
󰆓 +󰆓+󰆓
󰆓
󰆓 and 󰆓󰆓

󰆓󰆓 =
󰆒󰆒

󰆓󰆓 +󰆒󰆒

󰆓󰆓 +󰆒󰆒

󰆓󰆓 .
Writing the numeric value of the reaction rate r2 explicitly and reorganizing (6), we
can obtain
Remote Sens. 2022, 14, 3866 4 of 12
[][]., (7
)
where = 6.1 10298. and =󰆓󰆓
󰆓
󰆓
󰆓󰆓
󰆓󰆓
󰆓󰆓 , (= 0, <󰆒,󰆒󰆒 <).
Note that the coefficient ε depends on r2 and, therefore, can vary. For example, refs.
[29,31] utilized r2 = 1.2 × 10−27 after the work in [42]. The other examples of r2 applied in
previous studies include 2.7 × 10−34∙3002.4 [40], 1.4 × 10−34∙3002.4 [35], and 1.5 × 10−34∙3002.4 [43].
Despite the differences, all the studies were in consensus that ~..
Figure 1. Night-time zonal mean quantities averaged between 70°N and 90°N and over the period
of solar longitudes Ls = 265°–320°: (a) O, O3, H, O2, CO2, N2, T from MCD; (b) OHv = 1,…,9, calculated
with (1) (solid lines) and estimated with (7) (dashed lines); (c) volume emissions from (1) and (7)
(solid and dashed lines, respectively) for vibrational transitions 1–0 (blue), 2–1 (green), and 2–0 (red).
2.1. Peak Concentration of the Excited Hydroxyl Layer and Its Altitude
We now can derive an expression for the peak concentration of the hydroxyl layer
OH* and its altitude. For that, we exclude air density M from (7) using the ideal gas law:
[].[], (8
)
where the notation =
is used, p is pressure, and is the Boltzmann constant.
Differentiating (8) by pressure and equating the result to zero gives the pressure at
the local maximum of OH* concentration:
1
3.4 ln
 ln[]
 1
ln .
[], (9
)
Substituting (9) into (8), we obtain the value of the maximum concentration of the
excited hydroxyl:
[].[]
3.4 ln
 ln[]
 .[]
ln .
[], (10
)
It is seen from (9) and (10) that the peak concentration of OH* and its height are ex-
plicitly determined by vertical profiles of temperature, the concentration of atomic oxy-
gen, and the coefficient , which encompasses photochemical parameters. Note that the
derivations above are valid only within a thin layer near the peak of the OH* layer because
several assumptions were utilized that are only valid in this region.
2.2. Variations of the Excited Hydroxyl Layer
The hydroxyl layer is extremely variable. Therefore, it is desirable to link its relative
variations to those of the observable background quantities. For that, we decompose the
atomic oxygen number density, temperature, and air number density into the mean
Remote Sens. 2022, 14, 3866 5 of 12
([]
,,
) and deviations ([]󰆒,󰆒,), where the bar denotes an appropriate (spatial, tem-
poral, or both) averaging, and substitute them into (7):
[]=[]
+[](+󰆒).[]
+[]. (11
)
Temperature variations 󰆒/ are small, at least on Mars and other planets of the ter-
restrial group. This allows one to apply the Taylor series expansion to the term with tem-
perature in (11). Cross-multiplying all terms yields
[][]
.[]
+[]
.[]󰆒+[]󰆒.[]
2.4[]
󰆒.[]
+[]󰆒.[]󰆒
2.4[]
󰆒.[]󰆒2.4[]󰆒󰆒.[]
2.4[]󰆒󰆒.[]󰆒.
(12
)
The excited hydroxyl concentration for a given vibrational number can be written in
a more compact form:
[][]
+[]󰆒+[]󰆒+[]󰆒+[]󰆒󰆒+[]󰆒󰆒
+[]󰆒󰆒+  , (13
)
where the following notations are used: []
=.[]
[]
,[]󰆒=
.[]
[]󰆒,[]󰆒=.[]󰆒[]
,[]󰆒=2.4󰆒.[]
[]
,[]󰆒󰆒=
.[]󰆒[]󰆒,[]󰆒󰆒=2.4󰆒.[]
[]󰆒,[]󰆒󰆒=2.4󰆒.[]󰆒[]
.
Hence, relative variations of OH* concentration due to linear parts (RV) can be ex-
pressed in terms of the relative variations of temperature, atomic oxygen, and concentra-
tion of air: 󰆒[]
󰆒
[]
=2.4 󰆒
,
󰆒[]
󰆒
[]
=[]󰆒
[]
,

󰆒[]
󰆒
[]
=[]󰆒
[]
.
(14
)
The relative variations of the concentration due to second momenta (RV) are

󰆒󰆒 []
󰆒󰆒
[]
=2.4 󰆒[]󰆒
[]
,

󰆒󰆒 []
󰆒󰆒
[]
=[]󰆒[]󰆒
[]
[]
,

󰆒󰆒 []
󰆒󰆒
[]
=2.4 󰆒[]󰆒
[]
.
(15
)
In the derivation of (14) and (15), namely, in handling the terms with air number
density, we assumed that variations of the height of the OH* layer do not exceed the air
density scale height. Therefore, the derived equations are only valid when the displace-
ments of the OH* layer from the average altitude do not exceed the air density scale
height. In the terrestrial atmosphere, this condition is fulfilled for day-to-day, intra-sea-
sonal, gravity wave-induced variations and for annual cycles at latitudes where height
deviations of the OH* layer are relatively small. Similar care should be taken when (14)
and (15) are applied on Mars.
3. Calculations and Discussion
In this section, we test the applicability of the derived formulae. They contain photo-
chemical parameters in the most general form. In particular, they assume multi-quantum
relaxation for quenching and spontaneous emission processes, where transitions occur
from all vibrational levels above to all levels below. To date, not all multi-quantum
quenching coefficients for carbon dioxide and molecular nitrogen are known. Only the
Remote Sens. 2022, 14, 3866 6 of 12
rates for the so-called collisional cascade quenching [33], where transitions take place to
one level below, have been provided in the literature. The most recent update for these
coefficients was presented by [28,29] for quenching by carbon dioxide and molecular ni-
trogen, respectively. We adopted these values in our calculations. Namely, we used the
diagonal matrix for Avv and Gvv for transitions v v 1 with values of [28,29] and as-
signed the non-diagonal terms for other transitions to zero.
The input profiles of O, O3, H, O2, CO2, and N2 concentrations, and temperature, were
taken from the Mars Climate Database (MCD), which is based on simulations with the
Laboratoire de Météorologie Dynamique General Circulation Model (LMD-GCM) [44,45].
The MCD contains distributions of minor gases in the Martian atmosphere, including
ozone [46], which is directly involved in OH* production; water vapor [47], which is the
principal source of odd-hydrogens (H, OH, HO2); and variations of other long-lived spe-
cies (carbon dioxide and molecular nitrogen) involved in quenching processes [48,49].
Figure 1a presents the input profiles of night-time O, O3, H, O2, CO2, and temperature
T from the MCD averaged zonally between 70°N and 90°N and over the interval of solar
longitudes Ls = 265°320°. The averaging over this region and time period has been per-
formed in order to provide a colocation with the observations [24]. The results of calcula-
tions for [OH*] using the general Formula (1) and approximated by (7) for OHv = 1,…,9 are
shown in Figure 1b with solid and dashed lines, correspondingly.
The results illustrate good agreement between OH* concentrations and peak alti-
tudes calculated with the full model (1) and the simplified formula (7). The best agreement
occurs near the peaks at ~48–53 km. The differences below and above the maxima can be
partially explained by deviations of ozone from photochemical equilibrium in the polar
night region, where the ozone lifetime is prolonged under the condition of permanent
night and downward transport of atomic oxygen [50].
The vertical separation of the hydroxyl layer depending on vibrational numbers is
well-known in Earths atmosphere, e.g., [26,51]. It cannot be explained from (9) since v
does not depend on p. This is the result of omitting quenching by atomic oxygen in the
loss term for excited hydroxyl. The inclusion of this term produces a weak vertical sepa-
ration by vibrational numbers (solid lines). Vertical distances between layers correspond-
ing to different vibrational numbers are expected to be smaller on Mars than on Earth, as
was found by [24]. This is because the atomic oxygen quenching, which is responsible for
separation, is comparable with that of molecular oxygen near the Earth mesopause but is
negligible compared to the CO2 quenching in the Martian atmosphere.
The increase in excited hydroxyl concentration with decreasing vibrational number
was found from observations and modeling for the Earth's atmosphere [26,28,32,33,51]
and from modeling results for the Martian atmosphere [31]. To explain this fact, let us
consider Equation (6). Direct population from the reaction of ozone with atomic hydrogen
(first term in the numerator) is a slower process than population by quenching from the
upper vibrational levels (second term in the numerator). The second term in the numera-
tor (and therefore the whole numerator) increases with a decreasing vibrational number,
whereas the denominator can only decrease with a decline in the vibrational number.
Thus, the increase in the OH* concentration with a decreasing vibrational number be-
comes evident.
Volume emission is a measurable quantity that is proportional to the concentration
of OH*. We calculated it with the full formula (1) and approximated it by (7) (both assume
the photochemical equilibrium of excited hydroxyl) and plotted it in Figure 1c using solid
and dashed lines, respectively. The colors indicate the main vibrational transitions: 10
(blue), 21 (green), and 20 (red). The figure shows that the locations of peaks (at ~48–53
km) and the corresponding volume emissions are in good agreement with the observa-
tions of [24] in terms of shape and magnitude.
Equations derived in Section 2 provide some predictions and can be applied for anal-
ysis in the future, which we illustrate below. The terrestrial OH* airglow layer demon-
strates annual and semiannual variations [2,8,52,53]. Similar variations can be expected
Remote Sens. 2022, 14, 3866 7 of 12
from the Martian OH* due to seasonal changes in atomic oxygen, air number density, and
temperature.
Figure 2 shows time series of night-time one-month sliding averaged values at the
peak of the OHv = 2 layer calculated with (1) at middle (40° N) latitudes: (a) the concentra-
tion [OHv = 2], (b) the height of the peak, (c) the atomic oxygen concentration, and (d) tem-
perature. It is seen that the concentration and the height of the peak at the northern middle
latitude vary seasonally, with the maxim concentrations and lowest height occurring dur-
ing the first half of the year (Ls ≈ 0°180°). The amplitude of the annual height variation
on Mars is more than 20 km (Figure 2b), which by several times exceeds that near the
Earth mesopause (~5–10 km). The figures show a clear anticorrelation between the OHv = 2
number density and the height of the peak, as also follows from (8). Since volume emission
is linearly proportional to the hydroxyl concentration, this points to an anticorrelation be-
tween the emission and the height of the layer. A similar anticorrelation has also been
observed on Earth, e.g., [8,54,55].
Figure 2. Night-time mean one-month sliding averaged values at the peak of the OHv = 2 layer calcu-
lated with (1) at middle (40° N) latitudes: (a) concentration [OHv = 2], (b) the height of the peak, (c)
atomic oxygen concentration, and (d) temperature.
Figure 2a,c demonstrates a correlation between the concentrations of atomic oxygen
and excited hydroxyl. This correlation happens between Ls~210° and 340°, where the mi-
nor maximum of [OH*] coincides with the maximum of [O]. The correlation between the
air number density and the peak altitude is even more robust because the magnitude of
seasonal variations of the air density is larger than that of atomic oxygen. The effects of
atomic oxygen and air number densities on the OH* layer oppose each other. When the
OH* layer is low in summer, the air density is large, while the atomic oxygen concentra-
tion is small. The OH* layer moves higher in winter, and the air density decreases, but the
Remote Sens. 2022, 14, 3866 8 of 12
atomic oxygen concentration rises. In the Earths mesosphere at high and middle lati-
tudes, the behavior of the OH* layer is opposite: high altitude and low emission in sum-
mer, but a lower altitude and stronger emission in winter. This is because the main driver
for the OH* layer on Earth is atomic oxygen, which is transported downward in winter
and upward in summer. On Mars, the layer behavior is additionally determined by air
density variations. Seasonal changes in temperature play a minor role in the annual cycle
of OH* since it only varies by about 15 K over the year (Figure 2d).
In order to assess the sensitivity of the OH* layer to input parameters, we separately
calculated the contributions of relative variations of atomic oxygen, temperature, and air
density to variations of [OH*] or to the volume emission rate. We only considered the first
half of the year (Ls = 180°), during which displacements of the height of the layer did
not exceed the air density scale height (~10 km). Thus, the overbar in (14) and (15) denotes
a semiannual averaging, and primes are for deviations from the semiannual mean. As in
Figure 2, we only considered night-time values at 40° N, which were smoothed with the
one-month moving window averaging.
The results are plotted in Figure 3, with contributions from (14) and (15) shown by
solid and dashed lines, respectively. The figure illustrates our notion above that tempera-
ture (red lines) plays a minor role in the hydroxyl layer variability. The main contribution
comes from variations of atomic oxygen and the ambient air concentration acting in op-
posite phases. The first peak of [OH*] at Ls~40°50° (Figure 2a) is primarily determined by
the growth of the air number density (blue line) and, to a much lesser degree, by the de-
clining temperature (red line, see also Figure 2d). The secondary peak of [OH*] around
Ls~150° is mainly caused by the increase in the atomic oxygen concentration (green line),
whereas the declining air density and rising temperature act in the opposite direction. The
variations due to second momenta (dashed lines) are much weaker (do not exceed 10%).
Figure 3. Relative variations calculated over the first half of the Martian year with (14) (solid lines)
and (15) (dashed lines) at 40°N.
4. Conclusions
We presented the derivation of the simplified formulae relating the height of the peak
of the excited hydroxyl layer, its displacements, and the strength of emission with values
Remote Sens. 2022, 14, 3866 9 of 12
that can be observed in the Martian atmosphere at night-time. The assumptions used in
the derivation and relevant for Mars conditions include (1) the photochemical equilibrium
of ozone near the peak of the layer and (2) that total quenching by carbon dioxide, molec-
ular oxygen, and molecular nitrogen is greater than quenching by atomic oxygen and
spontaneous emission.
Under these approximations, the night-time concentration of OH* near the peak is
directly proportional to the concentration of atomic oxygen and pressure and inversely to
the power of temperature. Since pressure drops with altitude, the hydroxyl emission, the
major part of which is produced in the vicinity of the peak, anticorrelates with the height
of the OH* layer.
Calculations using input parameters from the Mars Climate Database demonstrate
annual variations of the OH* layer at middle latitude (40°N) resulting from the seasonal
cycle of temperature, air number density, and atomic oxygen. We illustrated how relative
variations of each of these quantities directly impact the relative variations of the concen-
tration of the hydroxyl layer.
The presented approach and simplified formulae can be applied for the analysis and
interpretation of future observations of hydroxyl emission on Mars. Coupled with obser-
vations of temperature and atomic oxygen (or ozone), airglow measurements can reveal
additional information about the Martian atmosphere’s dynamics and composition.
Finally, we should note that Equations (9) and (10) introduce the possibility of infer-
ring the altitudes of the OH* peak, the concentration at peak, the atomic oxygen concen-
tration at peak, and the ground-state hydroxyl concentration (which is the key constituent
for resolving the problem of the Martian (CO2) atmosphere stability due to catalytic re-
combination [5658]), by surface-based or nadir observations of emissions from two vi-
brational transitions, accompanied by temperature observations from vibro-rotational
transitions following [59].
The strength and advantage of full models are that they seek to most fully encompass
processes occurring in a photochemical system. The advantage of an analytical approach
is that it allows inferring, under certain conditions and assumptions, simple relations
within this system. In this work, we did exactly the latter: derived simplified relations
between OH* peak height and density and observable parameters of emission. They
can/should motivate the development of future observations and help with interpreta-
tions when such observations become available.
Author Contributions: Conceptualization, M.G., D.S.S., A.S.M., G.R.S. and P.H.; methodology,
M.G., D.S.S., A.S.M., G.R.S. and P.H.; software, M.G., D.S.S., A.S.M., G.R.S. and P.H.; validation,
M.G., D.S.S., A.S.M., G.R.S. and P.H.; formal analysis, M.G., D.S.S., A.S.M., G.R.S. and P.H.; investi-
gation, M.G., D.S.S., A.S.M., G.R.S. and P.H.; writingoriginal draft preparation, M.G., D.S.S.,
A.S.M., G.R.S. and P.H.; writingreview and editing, M.G., D.S.S., A.S.M., G.R.S. and P.H.; visual-
ization, M.G., D.S.S., A.S.M., G.R.S. and P.H. All authors have read and agreed to the published
version of the manuscript.
Funding: This research was partially funded by Russian Science Foundation grant 20-72-00110.
Data Availability Statement: The MCD data were obtained from the webpage (http://www-
mars.lmd.jussieu.fr/, accessed on 12 October 2021). The results of calculations are stored at
https://doi.org/10.5281/zenodo.5558814.
Acknowledgments: The authors are grateful to Francois Forget and Jean-Paul Huot for creating the
open access Mars Climate Database and to all collaborators of LMD who worked on the database.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the
design of the study; in the collection, analyses, or interpretation of data; in the writing of the manu-
script, or in the decision to publish the results.
References
Remote Sens. 2022, 14, 3866 10 of 12
1. Lopez-Gonzalez, M.J.; Rodríguez, E.; Shepherd, G.G.; Sargoytchev, S.; Shepherd, M.G.; Aushev, V.M.; Brown, S.; García-Comas,
M.; Wiens, R.H. Tidal variations of O2 Atmospheric and OH(6-2) airglow and temperature at mid-latitudes from SATI observa-
tions. Ann. Geophys. 2005, 23, 3579–3590. https://doi.org/10.5194/angeo-23-3579-2005.
2. Xu, J.; Smith, A.K.; Jiang, G.; Gao, H.; Wei, Y.; Mlynczak, M.G.; Russell, J.M., III. Strong longitudinal variations in the OH
nightglow. Geophys. Res. Lett. 2010, 37, L21801. https://doi.org/10.1029/2010GL043972.
3. Buriti, R.A.; Takahashi, H.; Lima, L.M.; Medeiros, A.F. Equatorial planetary waves in the mesosphere observed by airglow
periodic oscillations. Adv. Space Res. 2005, 35, 2031–2036. https://doi.org/10.1016/j.asr.2005.07.012.
4. Lopez-Gonzalez, M.J.; Rodríguez, E.; García-Comas, M.; Costa, V.; Shepherd, M.G.; Shepherd, G.G.; Aushev, V.M.; Sargoytchev,
S. Climatology of planetary wave type oscillations with periods of 2-20 days derived from O2 atmospheric and OH(6-2) airglow
observations at mid-latitude with SATI. Ann. Geophys. 2009, 27, 3645–3662. https://doi.org/10.5194/angeo-27-3645-2009.
5. Taylor, M.J.; Espy, P.J.; Baker, D.J.; Sica, R.J.; Neal, P.C.; Pendleton, W.R., Jr. Simultaneous intensity, temperature and imaging
measurements of short period wave structure in the OH nightglow emission. Planet. Space Sci. 1991, 39, 1171–1188.
https://doi.org/10.1016/0032-0633(91)90169-B.
6. Shepherd, G.G.; Thuillier, G.; Cho, Y.-M.; Duboin, M.-L.; Evans, W.F.J.; Gault, W.A.; Hersom, C.; Kendall, D.J.W.; Lathuillère,
C.; Lowe, R.P.; et al. The Wind Imaging Interferometer (WINDII) on the Upper Atmosphere Research Satellite: A 20 year per-
spective. Rev. Geophys. 2012, 50, RG2007. https://doi.org/10.1029/2012RG000390.
7. Wachter, P.; Schmidt, C.; Wüst, S.; Bittner, M. Spatial gravity wave characteristics obtained from multiple OH(3-1) airglow
temperature time series. J. Atmos. Sol. Terr. Phys. 2015, 135, 192–201. https://doi.org/10.1016/j.jastp.2015.11.008.
8. Gao, H.; Xu, J.; Wu, Q. Seasonal and QBO variations in the OH nightglow emission observed by TIMED/SABER. J. Geophys. Res.
2010, 115, A06313. https://doi.org/10.1029/2009JA014641.
9. Shepherd, M.G.; Cho, Y.-M.; Shepherd, G.G.; Ward, W.; Drummond, J.R. Mesospheric temperature and atomic oxygen response
during the January 2009 major stratospheric warming. J. Geophys. Res. 2010, 115, A07318. https://doi.org/10.1029/2009JA015172.
10. Shepherd, M.G.; Meek, C.E.; Hocking, W.K.; Hall, C.M.; Partamies, N.; Sigernes, F.; Manson, A.H.; Ward, W.E. Multi-instrument
study of the mesosphere-lower thermosphere dynamics at 80°N during the major SSW in January 2019. J. Atmos. Sol. Terr. Phys.
2020, 210, 105427. https://doi.org/10.1016/j.jastp.2020.105427.
11. Bittner, M.; Offermann, D.; Graef, H.-H.; Donner, M.; Hamilton, K. An 18 year time series of OH rotational temperatures and
middle atmosphere decadal variations. J. Atmos. Sol. Terr. Phys. 2002, 64, 1147–1166. https://doi.org/10.1016/S1364-
6826(02)00065-2.
12. Espy, P.J.; Stegman, J.; Forkman, P.; Murtagh, D. Seasonal variation in the correlation of airglow temperature and emission rate,
Geophys. Res. Lett. 2007, 34, L17802. https://doi.org/10.1029/2007GL031034.
13. Pertsev, N.; Perminov, V. Response of the mesopause airglow to solar activity inferred from measurements at Zvenigorod,
Russia. Ann. Geophys. 2008, 26, 1049–1056. https://doi.org/10.5194/angeo-26-1049-2008.
14. Dalin, P.; Perminov, V.; Pertsev, N.; Romejko, V. Updated long-term trends in mesopause temperature, airglow emissions, and
noctilucent clouds. J. Geophys. Res. 2020, 125, e2019JD030814. https://doi.org/10.1029/2019JD030814.
15. Perminov, V.I.; Pertsev, N.N.; Dalin, P.A.; Zheleznov, Y.A.; Sukhodoev, V.A.; Orekhov, M.D. Seasonal and Long-Term Changes
in the Intensity of O2(b1Σ) and OH(X2Π) Airglow in the Mesopause Region. Geomagn. Aeron. 2021, 61, 589–599.
https://doi.org/10.1134/S0016793221040113.
16. Russell, J.P.; Ward, W.E.; Lowe, R.P.; Roble, R.G.; Shepherd, G.G.; Solheim, B. Atomic oxygen profiles (80 to 115 km) derived
from Wind Imaging Interferometer/Upper Atmospheric Research Satellite measurements of the hydroxyl and greenline air-
glow: Local timelatitude dependence. J. Geophys. Res. 2005, 110, D15305. https://doi.org/10.1029/2004JD005570.
17. Mlynczak, M.G.; Hunt, L.A.; Mast, J.C.; Marshall, B.T.; Russell, J.M., III; Smith, A.K.; Siskind, D.E.; Yee, J.-H.; Mertens, C.J.;
Martin-Torres, F.J.; et al. Atomic oxygen in the mesosphere and lower thermosphere derived from SABER: Algorithm theoreti-
cal basis and measurement uncertainty. J. Geophys. Res. 2013, 118, 5724–5735. https://doi.org/10.1002/jgrd.50401.
18. Mlynczak, M.G.; Hunt, L.A.; Marshall, B.T.; Mertens, C.J.; Marsh, D.R.; Smith, A.K.; Russell, J.M.; Siskind, D.E.; Gordley, L.L.
Atomic hydrogen in the mesopause region derived from SABER: Algorithm theoretical basis, measurement uncertainty, and
results, J. Geophys. Res. 2014, 119, 3516–3526. https://doi.org/10.1002/2013JD021263.
19. Piccioni, G.; Drossart, P.; Zasova, L.; Migliorini, A.; Gérard, J.-C.; Mills, F.P.; Shakun, A.; García Muñoz, A.; Ignatiev, N.; Grassi,
D.; et al. First detection of hydroxyl in the atmosphere of Venus. Astron. Astrophys. 2008, 483, L29–L33.
https://doi.org/10.1051/0004-6361:200809761.
20. Gérard, J.-C.; Soret, L.; Saglam, A.; Piccioni, G.; Drossart, P. The distributions of the OH Meinel and O2(a1 X3Σ) nightglow
emissions in the Venus mesosphere based on VIRTIS observations. Adv. Space Res. 2010, 45, 1268–1275.
https://doi.org/10.1016/j.asr.2010.01.022.
21. Soret, L.; Gérard, J.-C.; Piccioni, G.; Drossart, P. Venus OH nightglow distribution based on VIRTIS limb observations from
Venus Express. Geophys. Res. Lett. 2010, 37, L06805. https://doi.org/10.1029/2010GL042377.
22. Migliorini, A.; Piccioni, G.; Cardesín Moinelo, A.; Drossart, P. Hydroxyl airglow on Venus in comparison with Earth. Planet.
Space Sci. 2011, 59, 974–980. https://doi.org/10.1016/j.pss.2010.05.004.
23. Migliorini, A.; Piccioni, G.; Capaccioni, F.; Filacchione, G.; Tosi, F.; Gérard, J.C. Comparative analysis of airglow emissions in
terres-trial planets, observed with VIRTIS-M instruments on board Rosetta and Venus Express. Icarus 2013, 226, 1115–1127.
https://doi.org/10.1016/j.icarus.2013.07.027.
Remote Sens. 2022, 14, 3866 11 of 12
24. Clancy, R.T.; Sandor, B.J.; García-Muñoz, A.; Lefèvre, F.; Smith, M.D.; Wolff, M.J.; Montmessin, F.; Murchie, S.L.; Nair, H. First
detection of Mars atmospheric hydroxyl: CRISM Near-IR measurement versus LMD GCM simulation of OH Meinel band emis-
sion in the Mars polar winter atmosphere. Icarus 2013, 226, 272–281. https://doi.org/10.1016/j.icarus.2013.05.035.
25. Burkholder, J.B.; Sander, S.P.; Abbatt, J.; Barker, J.R.; Cappa, C.; Crounse, J.D.; Dibble, T.S.; Huie, R.E.; Kolb, C.E.; Kurylo, M.J.;
et al. Chemical Kinetics and Photochemical Data for Use in Atmospheric Studies; Evaluation No. 19; JPL Publication 19-5; Jet Propul-
sion Laboratory: Pasadena, CA, USA, 2020. Available online: http://jpldataeval.jpl.nasa.gov (accessed on 12 October 2021).
26. Adler-Golden, S. Kinetic parameters for OH nightglow modeling consistent with recent laboratory measurements. J. Geophys.
Res. 1997, 102, 19969–19976. https://doi.org/10.1029/97JA01622.
27. Caridade, P.J.S.B.; Horta, J.-Z.J.; Varandas, A.J.C. Implications of the O + OH reaction in hydroxyl nightglow modeling. Atmos.
Chem. Phys. 2013, 13, 1–13. https://doi.org/10.5194/acp-13-1-2013.
28. Makhlouf, U.B.; Picard, R.H.; Winick, J.R. Photochemical-dynamical modeling of the measured response of airglow to gravity
waves. 1. Basic model for OH airglow. J. Geophys. Res. 1995, 100, 11289–11311. https://doi.org/10.1029/94JD03327.
29. Krasnopolsky, V.A. Nighttime photochemical model and night airglow on Venus. Planet. Space Sci. 2013, 85, 78–88.
https://doi.org/10.1016/j.pss.2013.05.022.
30. Xu, J.; Gao, H.; Smith, A.K.; Zhu, Y. Using TIMED/SABER nightglow observations to investigate hydroxyl emission mechanisms
in the mesopause region. J. Geophys. Res. 2012, 117, D02301. https://doi.org/10.1029/2011JD016342.
31. García-Muñoz, A.; McConnell, J.C.; McDade, I.C.; Melo, S.M.L. Airglow on Mars: Some model expectations for the OH Meinel
bands and the O2 IR atmospheric band. Icarus 2005, 176, 75–95. https://doi.org/10.1016/j.icarus.2005.01.006.
32. Llewellyn, E.J.; Long, B.H.; Solheim, B.H. The quenching of OH* in the atmosphere. Planet. Space Sci. 1978, 26, 525–531.
https://doi.org/10.1016/0032-0633(78)90043-0.
33. McDade, I.C.; Llewellyn, E.J. Kinetic parameters related to sources and sinks of vibrationally excited OH in the nightglow, J.
Geophys. Res. 1987, 92, 7643–7650. https://doi.org/10.1029/JA092iA07p07643.
34. Meriwether, J.W., Jr. A review of the photochemistry of selected nightglow emissions from the mesopause. J. Geophys. Res. 1989,
94, 14629–14646. https://doi.org/10.1029/JD094iD12p14629.
35. Krasnopolsky, V.A.; Lefèvre, F. Chemistry of the atmospheres of Mars, Venus, and Titan. In Comparative Climatology of Terrestrial
Planets, 1st ed.; Mackwell, S.J., Simon-Miller, A.A., Eds.; University of Arizona: Tucson, AZ, USA, 2013; pp. 231–275.
https://doi.org/10.2458/azu_uapress_9780816530595-ch11.
36. Nair, H.; Allen, M.; Anbar, A.D.; Yung, Y.L.; Clancy, R.T. A Photochemical Model of the Martian Atmosphere. Icarus 1994, 111,
124–150. https://doi.org/10.1006/icar.1994.1137.
37. Dodd, J.A.; Lipson, S.J.; Blumberg, W.A.M. Formation and vibrational relaxation of OH(X2Πi, v) by O2 and CO2. J. Chem. Phys.
1991, 95, 5752–5762. https://doi.org/10.1063/1.461597.
38. Chalamala, B.R.; Copeland, R.A. Collision dynamics of OH(X2Π, v = 9). J. Chem. Phys. 1993, 99, 5807–5811.
https://doi.org/10.1063/1.465932.
39. Soret, L.; Gérard, J.-C.; Piccioni, G.; Drossart, P. The OH Venus nightglow spectrum: Intensity and vibrational composition from
VIRTIS Venus Express observations. Planet. Space Sci. 2012, 73, 387–396. https://doi.org/10.1016/j.pss.2012.07.027.
40. Krasnopolsky, V.A. Photochemistry of the martian atmosphere: Seasonal, latitudinal, and diurnal variations. Icarus 2006, 185,
153–170. https://doi.org/10.1016/j.icarus.2006.06.003.
41. Krasnopolsky, V.A. Solar activity variations of thermospheric temperatures on Mars and a problem of CO in the lower atmos-
phere. Icarus 2010, 207, 638–647. https://doi.org/10.1016/j.icarus.2009.12.036.
42. Lindner, B.L. Ozone on Mars: The effects of clouds and airborne dust. Planet. Space Sci. 1988, 36, 125–144.
https://doi.org/10.1016/0032-0633(88)90049-9.
43. Lefèvre, F.; Lebonnois, S.; Montmessin, F.; Forget, F. Three-dimensional modeling of ozone on Mars. J. Geophys. Res. 2004, 109,
E07004. https://doi.org/10.1029/2004JE002268.
44. Forget, F.; Hourdin, F.; Fournier, R.; Hourdin, C.; Talagrand, O.; Collins, M.; Lewis, S.R.; Read, P.L.; Huot, J.-P. Improved general
circulation models of the Martian atmosphere from the surface to above 80 km. J. Geophys. Res. 1999, 104, 24155–24176.
https://doi.org/10.1029/1999JE001025.
45. Millour, E.; Forget, F.; Spiga, A.; Vals, M.; Zakharov, V.; Montabone, L.; Lefèvre, F.; Montmessin, F.; Chaufray, J.-Y.; López-
Valverde, M.A.; et al. The Mars Climate Database (Version 5.3). In Proceedings of the Scientific Workshop: From Mars Express
to ExoMars, ESAC, Madrid, Spain, 2728 February 2018. Available online: https://ui.adsabs.harvard.edu/link_gate-
way/2018fmee.confE..68M/PUB_PDF (accessed on 12 October 2021).
46. Lefèvre, F.; Bertaux, J.-L.; Clancy, R.T.; Encrenaz, T.; Fast, K.; Forget, F.; Lebonnois, S.; Montmessin, F.; Perrier, S. Heterogeneous
chemistry in the atmosphere of Mars. Nature 2008, 454, 971–975. https://doi.org/10.1038/nature07116.
47. Navarro, T.; Madeleine, J.-B.; Forget, F.; Spiga, A.; Millour, E.; Montmessin, F.; Määttänen, A. Global climate modeling of the
Martian water cycle with improved microphysics and radiatively active water ice clouds. J. Geophys. Res. 2014, 119, 1479–1495.
https://doi.org/10.1002/2013JE004550.
48. Forget, F.; Hourdin, F.; Talagrand, O. CO2 Snowfall on Mars: Simulation with a General Circulation Model. Icarus 1998, 131,
302–316. https://doi.org/10.1006/icar.1997.5874.
49. Forget, F.; Millour, E.; Montabone, L.; Lefevre, F. Non condensable gas enrichment and depletion in the Martian polar regions.
In Proceedings of the 3rd Workshop Mars Atmosphere: Modeling and Observations, Williamsburg, VA, USA, 1013 November
2008.
Remote Sens. 2022, 14, 3866 12 of 12
50. Shaposhnikov, D.S.; Medvedev, A.S.; Rodin, A.V.; Hartogh, P. Seasonal water “pump” in theatmosphere of mars: Vertical
transport to the thermosphere. Geophys. Res. Lett. 2019, 46, 4161–4169. https://doi.org/10.1029/2019GL082839.
51. Swenson, G.R.; Gardner, C.S. Analytical models for the resposes of the mesospheric OH* and Na layers to atmospheric gravity
waves. J. Geophys. Res. 1998, 103, 6271–6294. https://doi.org/10.1029/97JD02985.
52. Liu, G.; Shepherd, G.G.; Roble, R.G. Seasonal variations of the nighttime O(1S) and OH airglow emission rates at mid-to-high
latitudes in the context of the large-scale circulation. J. Geophys. Res. 2008, 113, A06302. https://doi.org/10.1029/2007JA012854.
53. Marsh, D.R.; Smith, A.K.; Mlynczak, M.G.; Russell, J.M., III. SABER observations of the OH Meinel airglow variability near the
mesopause. J. Geophys. Res. 2006, 111, A10S05. https://doi.org/10.1029/2005JA011451.
54. Liu, G.; Shepherd, G.G. An empirical model for the altitude of the OH nightglow emission. Geophys. Res. Lett. 2006, 33, L09805.
https://doi.org/10.1029/2005GL025297.
55. Mulligan, F.G.; Dyrland, M.E.; Sigernes, F.; Deehr, C.S. Inferring hydroxyl layer peak heights from ground-based measurements
of OH(62) band integrated emission rate at Longyearbyen (78°N, 16°E). Ann. Geophys. 2009, 27, 4197–4205.
https://doi.org/10.5194/angeo-27-4197-2009.
56. McElroy, M.B.; Donahue, T.M. Stability of the Martian Atmosphere. Science 1972, 177, 986–988. https://doi.org/10.1126/sci-
ence.177.4053.986.
57. Parkinson, T.D.; Hunten, D.M. Spectroscopy and aeronomy of O2 on Mars. J. Atmos. Sci. 1972, 29, 1380–1390.
https://doi.org/10.1175/1520-0469(1972)029<1380:SAAOOO>2.0.CO;2.
58. Olsen, K.S.; Lefèvre, F.; Montmessin, F.; Fedorova, A.A.; Trokhimovskiy, A.; Baggio, L.; Korablev, O.; Alday, J.; Wilson, C.F.;
Forget, F.; et al. The vertical structure of CO in the Martian atmosphere from the ExoMars Trace Gas Orbiter. Nat. Geosci. 2021,
14, 67–71. https://doi.org/10.1038/s41561-020-00678-w.
59. Conway, S. Methods for Deriving Temperature Profiles of Mars from OH Meinel Airglow Observations. Ph.D. Thesis, York
University, Toronto, ON, Canada, March 2012.
... We assume that the excited hydroxyl is in a photochemical equilibrium during nighttime [43]. This assumption enables us to explicitly express the concentration of hydroxyl at all excitation levels [OH v ] in the following form (see [44], Equation (1)): ...
... The model outlined above has been described in detail and compared with the available observations [39] in our previous studies [44,57]. It reproduces the values and shape of the CRISM observations for transitions 1-0 and 2-0 and shows~30% lower values near the peak for transition 2-1. ...
... It reproduces the values and shape of the CRISM observations for transitions 1-0 and 2-0 and shows~30% lower values near the peak for transition 2-1. The latter is still within the CRISM's uncertainty (see Figure 1c in [44] and Figure 7 in [39]). For transition 1-0, this result is better than the simulations with the Laboratoire de Météorologie Dynamique General Circulation Model (LMD GCM) and with the OH* models of Krasnopolsky [58] and García-Muñoz [43]. ...
Article
Full-text available
Monitoring excited hydroxyl (OH*) airglow is broadly used for characterizing the state and dynamics of the terrestrial atmosphere. Recently, the existence of excited hydroxyl was confirmed using satellite observations in the Martian atmosphere. The location and timing of its detection on Mars were restricted to a winter season at the north pole. We present three-dimensional global simulations of excited hydroxyl over a Martian year. The predicted spatio-temporal distribution of the OH* can provide guidance for future observations, namely by indicating where and when the airglow is likely to be detected.
Article
Full-text available
Carbon monoxide (CO) is the main product of CO2 photolysis in the Martian atmosphere. Production of CO is balanced by its loss reaction with OH, which recycles CO into CO2. CO is therefore a sensitive tracer of the OH-catalysed chemistry that contributes to the stability of CO2 in the atmosphere of Mars. To date, CO has been measured only in terms of vertically integrated column abundances, and the upper atmosphere, where CO is produced, is largely unconstrained by observations. Here we report vertical profiles of CO from 10 to 120 km, and from a broad range of latitudes, inferred from the Atmospheric Chemistry Suite on board the ExoMars Trace Gas Orbiter. At solar longitudes 164–190°, we observe an equatorial CO mixing ratio of ~1,000 ppmv (10–80 km), increasing towards the polar regions to more than 3,000 ppmv under the influence of downward transport of CO from the upper atmosphere, providing a view of the Hadley cell circulation at Mars’s equinox. Observations also cover the 2018 global dust storm, during which we observe a prominent depletion in the CO mixing ratio up to 100 km. This is indicative of increased CO oxidation in a context of unusually large high-altitude water vapour, boosting OH abundance.
Article
Full-text available
We have updated long‐term trends in mesopause temperature, airglow emission intensities, and noctilucent clouds (NLCs) based on ground‐based observations conducted in the Moscow region (Russia). Trends in mesopause temperature and airglow emissions have been derived for the period 2000–2018 (19 years), and long‐term trends in NLC characteristics have been obtained for 1968–2018 (51 years). Trends in airglow emissions have been estimated separately for winter and summer seasons. There are statistically significant negative trends in molecular oxygen O2 A(0‐1) and hydroxyl OH(6‐2) emission intensities in both winter and summer. Responses of the airglow intensities to solar activity have been found to be significantly positive, with winter ones being 1.5–2 times greater than summer ones. There is a rather strong and statistically significant cooling of the summer mesopause at a rate of −2.4±2.3 K/decade, whereas the winter mesopause demonstrates a small and statistically insignificant cooling of −0.4±2.2 K/decade. The response of the mesopause temperature to solar activity is positive and about 2 times greater in winter than in summer. We consider long‐term trends in the summer mesopause temperature and NLC for the same time period and same geographical region. Secular trends in NLC parameters have been found to be small and statistically insignificant as observed from midlatitudes. NLCs demonstrate statistically significant negative response to solar activity. Ozone forcing produces minor effects on both airglow and NLC characteristics. The observed small secular NLC trends contradict large modeled NLC trends recently obtained by Lübken et al. (2018, https://doi.org/10.1029/2018GL077719).
Article
Full-text available
We present results of simulations with the Max Planck Institute Martian general circulation model implementing a hydrological cycle scheme. The simulations reveal a seasonal water “pump” mechanism responsible for the upward transport of water vapor. This mechanism occurs in high latitudes above 60° of the southern hemisphere at perihelion, when the upward branch of the meridional circulation is particularly strong. A combination of the mean vertical flux with variations induced by solar tides facilitates penetration of water across the “bottleneck” at approximately 60 km. The meridional circulation then transports water across the globe to the northern hemisphere. Since the intensity of the meridional cell is tightly controlled by airborne dust, the water abundance in the thermosphere strongly increases during dust storms.
Article
Full-text available
The hydroxyl nightglow has been examined anew using calculated rate constants for the key reactive and inelastic O + OH( v ') quenching processes. These constants have been obtained from quasiclassical trajectories run on the adiabatic ab initio-based double many-body expansion-IV potential energy surface for the ground state of the hydroperoxil radical. Significant differences in the vertical profiles of vibrationally excited hydroxyl radicals are obtained relative to the ones predicted by Adler-Golden (1997) when employing an O + OH( v ') effective rate constant chosen to be twice the experimental value for quenching of OH( v ' = 1). At an altitude of 90 km, such deviations range from ~ 80% for v ' = 1 to only a few percent for v ' = 9. Other mechanisms reported in the literature have also been utilized, in particular those that loosely yield lower and upper limits in the results, namely sudden-death and collisional cascade. Finally, the validity of the steady-state hypothesis is analysed through comparison with results obtained via numerical integration of the master equations.
Book
Full-text available
1. INTRODUCTION Energies of the solar UV photons and energetic particles may be sufficient to break chemical bonds in atmospheric species and form new molecules, atoms, radicals, ions, and free electrons. These products initiate chemical reactions that further complicate the atmospheric composition, which is also significantly affected by dynamics and transport processes. Photochemical products may be tracers of photochemistry and dynamics in the atmosphere and change its thermal balance. Gas exchange between the atmosphere, space, and solid planet also determines the properties of the atmosphere. Studies of the atmospheric chemical composition and its variations are therefore essential for all aspects of atmospheric science, and many instruments on spacecraft missions to the planets are designed for this task. Interpretation of the observations requires photochemical models that could adequately simulate photochemical and transport processes. One-dimensional steady-state global-mean self-consistent models are a basic type of photochemical modeling. These models simulate altitude variations of species in an atmosphere. Vertical transport in the model is described by eddy diffusion coefficient K(z). The solar UV spectrum, absorption cross sections, reaction rate coefficients, eddy and molecular diffusion coefficients K(z) and D(z), and temperature profile T(z) are the input data for the model. The model provides a set of continuity equations for each species in a spherical atmosphere 1 2 2 1 1 1 1 r d r dr P nL K D dn dr n K D H T dT dr H T dT a Φ Φ ( ) = − = − + ( ) −       + + + + α d dr             Here r is radius, Φ and n are the flux and number density of species, P and L are the production and loss of species in chemical reactions, α is the thermal diffusion factor, Ha and H are the mean and species scale heights, and T is the temperature. Substitution of the second equation in the first results in an ordinary second-order nonlinear differential equation. Finite-difference analogs for these equations may be solved using methods described by Allen et al. (1981) and Krasnopolsky and Cruikshank (1999). Boundary conditions for the equations may be densities, fluxes, and velocities of the species. Fluxes and velocities are equal to zero at a chemically passive surface and at the exobase for molecules that do not escape. Requirements for the boundary conditions are discussed in Krasnopolsky (1995). Generally, the number of nonzero conditions should be equal to the number of chemical elements in the system. This type of photochemical modeling is a powerful tool for studying atmospheric chemical composition. For example, in the case of Titan, a model results in vertical profiles of 83 neutral species and 33 ions up to 1600 km using densities of N2 and CH4 near the surface (Krasnopolsky, 2009a, 2012c). Photochemical general circulation models (GCMs; see Forget and Lebonnois, this volume) present significant prog231 Chemistry of the Atmospheres of Mars, Venus, and Titan Vladimir A. Krasnopolsky Catholic University of America Franck Lefèvre Laboratoire Atmosphères et Observations Spatiales (LATMOS), Paris Observations and models for atmospheric chemical compositions of Mars, Venus, and Titan are briefly discussed. While the martian CO2-H2O photochemistry is comparatively simple, Mars’ obliquity, elliptic orbit, and rather thin atmosphere result in strong seasonal and latitudinal variations that are challenging in both observations and modeling. Venus’ atmosphere presents a large range of temperature and pressure conditions. The atmospheric chemistry involves species of seven elements, with sulfur and chlorine chemistries dominating up to 100 km. The atmosphere below 60 km became a subject of chemical kinetic modeling only recently. Photochemical modeling is especially impressive for Titan: Using the N2 and CH4 densities at the surface and temperature and eddy diffusion profiles, it is possible to calculate vertical profiles of numerous neutrals and ions throughout the atmosphere. The Cassini-Huygens observations have resulted in significant progress in understanding the chemistry of Titan’s atmosphere and ionosphere and provide an excellent basis for their modeling. Krasnopolsky V. A. and Lefèvre F. (2013) Chemistry of the atmospheres of Mars, Venus, and Titan. In Comparative Climatology of Terrestrial Planets (S. J. Mackwell et al., eds.), pp. 231–275. Univ. of Arizona, Tucson, DOI: 10.2458/azu_uapress_9780816530595-ch11. 232 Comparative Climatology of Terrestrial Planets ress in the study and reproduction of atmospheric properties under a great variety of conditions. These GCMs are the best models for studying variations of species with local time, latitude, season, and location. Some basic properties of the terrestrial planets and their atmospheres are listed in Table 1. They are discussed in detail in other chapters of this book. Below...
Article
Contemporaneous multi-instrument ground-based optical and meteor radar observations of OH and O2 airglow, temperature and neutral winds during the winter season of November 2018–February 2019 have been used to investigate the dynamics of the mesosphere/lower thermosphere (MLT) (80–100 km) region in the high Arctic at Svalbard, Norway (78°N, 16°E) and at Eureka, Canada (80°N, 274°E). Temperature observations by the MLS Aura satellite over the 20–90 km height range for the same period were also considered. The period is characterized by an unusual major sudden stratospheric warming (SSW) event that began with a displacement of the polar vortex around December 13, 2018 (DoY 347), followed by the vortex split on January 2, 2019 (DoY 367). The MLS Aura temperature observations outlined four periods of interest: 1) late November – early December (DoY 328–335) with cold temperature anomalies in the mesosphere and warm bursts at the stratopause; 2) December 13, 2018–January 2, 2019 (DoY 347–367) when the stratopause rapidly descended to 45 km and broke down, triggering the onset of a SSW; 3) January 3–20, 2019 (DoY 368–385) during the stratopause recovery phase, and 4) from January 21, 2019 (DoY 386) onward, with the formation of the elevated stratopause and gradual return to its pre-SSW height. Both airglow emissions, OH and O2 Atm, showed simultaneously significant depletion of the integrated emission rates (IER) and temperature decrease of the order of 50–60 K, indicating upwelling, depletion of the atomic oxygen and adiabatic cooling. These cold temperature anomalies were followed by enhancements in the observed airglow IERs on December 13, 2018 (DoY 347) and January 2, 2019 (DoY 367) accompanied by a decrease in the peak altitude of the OH layer suggesting down-welling and influx of atomic oxygen from the lower thermosphere. The observations revealed oscillations with periods of 4.5–7 days, 8–10 days, and 16–21 days consistent with previously reported planetary wave activity in the winter MLT region and during major stratospheric warming events. However, the results presented here show for the first time comparisons of the multi-instrument temperature observations at 78°N – 80°N, providing an indispensable tool in monitoring the dynamics over the polar cap in general, and in describing the regional dynamical response of the MLT region to major large-scale phenomena like stratospheric warmings, in particular.
Article
We present a new approach for the detection of gravity waves in OH-airglow observations at the measurement site Oberpfaffenhofen (11.27°E, 48.08°N), Germany. The measurements were performed at the German Remote Sensing Data Center (DFD) of the German Aerospace Center (DLR) during the period from February 4th, 2011 to July 6th, 2011. In this case study the observations were carried out by three identical Ground-based Infrared P-branch Spectrometers (GRIPS). These instruments provide OH(3-1) rotational temperature time series, which enable spatio-temporal investigations of gravity wave characteristics in the mesopause region. The instruments were aligned in such a way that their fields of view (FOV) formed an equilateral triangle in the OH-emission layer at a height of 87. km. The Harmonic Analysis is applied in order to identify joint temperature oscillations in the three individual datasets. Dependent on the specific gravity wave activity in a single night, it is possible to detect up to four different wave patterns with this method. The values obtained for the waves' periods and phases are then used to derive further parameters, such as horizontal wavelength, phase velocity and the direction of propagation.We identify systematic relationships between periods and amplitudes as well as between periods and horizontal wavelengths. A predominant propagation direction towards the East and North-North-East characterizes the waves during the observation period. There are also indications of seasonal effects in the temporal development of the horizontal wavelength and the phase velocity. During late winter and early spring the derived horizontal wavelengths and the phase velocities are smaller than in the subsequent period from early April to July 2011.
Article
The photochemical model for the Venus nighttime atmosphere and night airglow (Krasnopolsky, 2010, Icarus 207, 17-27) has been revised to account for the SPICAV detection of the nighttime ozone layer and more detailed spectroscopy and morphology of the OH nightglow. Nighttime chemistry on Venus is induced by fluxes of O, N, H, and Cl with mean hemispheric values of 3×1012, 1.2×109, 1010, and 1010 cm-2 s-1, respectively. These fluxes are proportional to column abundances of these species in the daytime atmosphere above 90 km, and this favors their validity. The model includes 86 reactions of 29 species. The calculated abundances of Cl2, ClO, and ClNO3 exceed a ppb level at 80-90 km, and perspectives of their detection are briefly discussed. Properties of the ozone layer in the model agree with those observed by SPICAV. An alternative model without the flux of Cl agrees with the observed O3 peak altitude and density but predicts an increase of ozone to 4×108 cm-3 at 80 km. Reactions H+O3 and O+HO2 that may excite the OH nightglow have equal column rates. However, the latter is shifted to 92-94 km, and the models agree better with the nightglow observations if O+HO2 does not contribute to the OH excitation. Schemes for quenching of the OH vibrational quanta by CO2 are chosen to fit the observed band distribution in the Δv=1 sequence at 2.9 μm. The models agree with all observational constraints for the mean nighttime atmosphere. Analytic relationships between the nightglow intensities, the ozone layer, and the input fluxes of atomic species are given. The model results are compared with those of three-dimensional models for the Venus thermosphere.
Article
Airglow emissions are optimal processes to investigate the chemistry and dynamics in planetary atmospheres. In this study, we focus on the O2 and OH airglow emissions in Venus, Earth, and Mars atmospheres, which are controlled by chemical reactions common to the three planets. By studying these phenomena on Venus, Earth, and Mars using similar instruments, we are able to derive information about their photochemistry and the physical conditions of the atmospheres, but also to constrain the dynamics responsible for transport of atomic oxygen, ozone and other minor species.