ArticlePDF Available

Secondary ice production during the break-up of freezing water drops on impact with ice particles

Authors:

Abstract and Figures

We provide the first dedicated laboratory study of collisions of supercooled water drops with ice particles as a secondary ice production mechanism. We experimentally investigated collisions of supercooled water drops (∼ 5 mm in diameter) with ice particles of a similar size (∼ 6 mm in diameter) placed on a glass slide at temperatures >-12 ∘C. Our results showed that secondary drops were generated during both the spreading and retraction phase of the supercooled water drop impact. The secondary drops generated during the spreading phase were emitted too fast to quantify. However, quantification of the secondary drops generated during the retraction phase with diameters >0.1 mm showed that 5–10 secondary drops formed per collision, with approximately 30 % of the secondary drops freezing over a temperature range between −4 and −12 ∘C. Our results suggest that this secondary ice production mechanism may be significant for ice formation in atmospheric clouds containing large supercooled drops and ice particles.
Content may be subject to copyright.
Atmos. Chem. Phys., 21, 18519–18530, 2021
https://doi.org/10.5194/acp-21-18519-2021
© Author(s) 2021. This work is distributed under
the Creative Commons Attribution 4.0 License.
Research article
Secondary ice production during the break-up of freezing
water drops on impact with ice particles
Rachel L. James1, Vaughan T. J. Phillips2, and Paul J. Connolly1
1Department of Earth and Environmental Sciences, The University of Manchester, Manchester, UK
2Department of Physical Geography and Ecosystem Science, Lund University, Lund, Sweden
Correspondence: Rachel L. James (rachel.james@manchester.ac.uk)
Received: 1 July 2021 Discussion started: 15 July 2021
Revised: 15 November 2021 Accepted: 18 November 2021 Published: 21 December 2021
Abstract. We provide the first dedicated laboratory study of collisions of supercooled water drops with ice par-
ticles as a secondary ice production mechanism. We experimentally investigated collisions of supercooled water
drops (5 mm in diameter) with ice particles of a similar size (6 mm in diameter) placed on a glass slide at
temperatures >12 C. Our results showed that secondary drops were generated during both the spreading and
retraction phase of the supercooled water drop impact. The secondary drops generated during the spreading phase
were emitted too fast to quantify. However, quantification of the secondary drops generated during the retraction
phase with diameters >0.1 mm showed that 5–10 secondary drops formed per collision, with approximately
30 % of the secondary drops freezing over a temperature range between 4 and 12 C. Our results suggest that
this secondary ice production mechanism may be significant for ice formation in atmospheric clouds containing
large supercooled drops and ice particles.
1 Introduction
Most surface rainfall events that occur across the globe
are associated with the ice phase within clouds in Earth’s
atmosphere (Field and Heymsfield, 2015), as are severe
weather events such as freezing rain, hail and thunderstorms
(Changnon, 2003; ˇ
cik et al., 2019; Elsom, 2001). There-
fore, understanding the processes which govern ice forma-
tion in clouds is crucial for determining their effects on both
climate and weather.
Where subzero temperatures are warmer than the homo-
geneous freezing point of 35 C, supercooled water drops
can heterogeneously freeze via a subset of aerosol particles
present in the atmosphere. This subset of aerosol particles,
called ice-nucleating particles (INPs), are relatively rare, and
while number concentrations of INPs vary in time and space,
they typically fall between 1 ×105to 1 L1at 10 C
(Kanji et al., 2017). Yet, observed ice particle concentrations
in mixed-phase clouds can be several orders of magnitude
higher than concentrations predicted from ice particles form-
ing due to INPs (e.g. Crawford et al., 2012; Lloyd et al.,
2015; Lasher-Trapp et al., 2016; Ladino et al., 2017). Ice can
also form at temperatures >35 C via secondary ice pro-
duction (SIP), where new ice particles are formed from pre-
existing ice particles. However, our understanding of ice for-
mation from SIP mechanisms is incomplete (e.g. see reviews
by Field et al., 2017; Korolev and Leisner, 2020), resulting in
poor representation of SIP mechanisms in numerical weather
prediction (NWP) models.
Observations within mixed-phase clouds often show ice
crystal number concentrations higher than the numbers of
INPs present in the atmosphere. For instance, ice particle
number concentrations exceeding 100 L1, in shallow con-
vection with a cloud top temperature no lower than 12 C,
have been observed over the UK (Crawford et al., 2012).
Furthermore, thin mixed-phase layer clouds have been ob-
served to continually generate snow (Westbrook and Illing-
worth, 2013). Conventional thinking would suggest that the
ice in mixed-phase layer clouds should fall out, leaving the
layer “depleted” of INPs; however, the observations clearly
show that ice continues to form in these clouds over time.
The rime splintering SIP mechanism has been successful
in predicting the glaciation of mixed-phase clouds in many
Published by Copernicus Publications on behalf of the European Geosciences Union.
18520 R. L. James et al.: Secondary ice formation during drop–ice collisions
cases, especially those involving a warm cloud base creating
sufficiently large cloud drops in the rime splintering temper-
ature region between 3 and 8C (e.g. Harris-Hobbs and
Cooper, 1987; Blyth and Latham, 1993, 1997; Phillips et al.,
2001, 2005; Crosier et al., 2011; Crawford et al., 2012; Tay-
lor et al., 2016; Huang et al., 2017). However, there are also
numerous cases where significant concentrations of ice ob-
served in clouds cannot be explained by the rime splintering
SIP mechanism. Hobbs and Rangno (1985) compiled tables
of aircraft observations from a wide range of cloud environ-
ments. They found that the maximum ice particle concentra-
tions were independent of the cloud top temperature but were
strongly dependent on the broadness of the supercooled drop
spectrum near the cloud top, with approximately half of the
clouds exhibiting ice enhancement.
Several SIP mechanisms have been identified and studied
both in the laboratory and theoretically, but only the rime
splintering SIP mechanism is widely implemented in NWP
models. Active between 3 and 8C, rime splintering oc-
curs when supercooled water drop diameters are <13 and
>24 µm (Hallett and Mossop, 1974; Mossop and Hallett,
1974; Mossop, 1978). Another SIP mechanism, the fragmen-
tation of freezing drops, has received a significant proportion
of laboratory-based SIP investigations. Fragmentation due to
freezing drizzle drops or raindrops can occur over a wider
temperature range between 0 and 32 C, but quantification
of ice fragment generation rates and temperature dependence
within these rates between laboratory studies varies signifi-
cantly (see Table 1 of Korolev and Leisner, 2020, for a sum-
mary). A range in diameters of freezing supercooled water
drops has also been investigated between laboratory studies
from 4 to 1000 µm (see Table 1 of Korolev and Leisner, 2020,
for a summary). While other SIP mechanisms exist (e.g. ice–
ice collisions and sublimation fragmentation), the attention
of laboratory studies has overwhelmingly focussed on the
SIP mechanisms of rime splintering and fragmentation due
to freezing drops. Furthermore, unidentified SIP mechanisms
may also exist.
In this paper, we present a SIP mechanism involving the
formation of secondary drops from the collision of a super-
cooled water drop with a larger ice particle. This SIP mecha-
nism has been investigated via a theoretical study by Phillips
et al. (2018), who referred to it as “mode 2”, as it involves
collisions of supercooled water drops with more massive ice
particles resulting in fragmentation of the supercooled water
drop. Ice contained in some of the secondary drops was as-
sumed to initiate freezing, yielding secondary ice fragments.
By contrast, “mode 1” involved either collisions of super-
cooled water drops with less massive ice particles resulting
in spherical freezing of the supercooled water drop or activa-
tion of immersed INPs, with a quasi-spherical outer ice shell
that fragments.
While there are no dedicated laboratory studies of this SIP
mechanism involving collisions of supercooled water drops
with more massive ice particles or the activation of INPs im-
mersed in them, there are laboratory studies that have indi-
rectly studied aspects of this process. For example, a sim-
ilar mechanism was alluded to by Latham and Warwicker
(1980) in their experimental investigation of charge transfer
during interactions between hailstones and supercooled wa-
ter drops. They observed that frost could occasionally be bro-
ken during impact, thus forming new ice particles. Although
this was an unwanted outcome of their experiments, it pro-
vided some hints of a potential SIP mechanism during the
interactions between ice particles and supercooled raindrops.
Later, Schremb et al. (2018) studied the fluid flow and solid-
ification of supercooled water drops on elevated ice targets,
briefly observing the formation of secondary drops from the
rim of the supercooled water drop during impact. However,
for both of these studies no quantification of the secondary
drops was made.
In this paper, we describe a set of experiments performed
at the University of Manchester to determine the freezing
fraction of secondary drops (8) formed in the splash dur-
ing the collision of a 5 mm diameter supercooled raindrop on
a 6 mm diameter ice particle, providing the first laboratory
quantification of this SIP mechanism. This freezing fraction
(8) is the ratio of secondary drops that freeze to all such
drops emitted. The experimental setup is described in Sect. 2.
The results are presented in Sect. 3, and the discussion is in
Sect. 4. Finally, the conclusions are given in Sect. 5.
2 Experimental setup
A schematic of the experimental setup is shown in Fig. 1. The
setup was purpose-built to study the impact of a supercooled
water drop on an ice particle. For this study, we used two con-
figurations of the experimental setup. The first configuration
was used to study the drop impact with a high-speed camera
(Chronos 1.4, Kron Technologies Inc.) equipped with a mi-
croscopic lens (Kron Technologies Inc.) and a 0.5×Barlow
lens (Kron Technologies Inc.) in a side-on view. The second
configuration was used to quantify the fraction of secondary
drops that froze after impact with the ice particle using two
Raspberry Pi Camera modules (Raspberry Pi Camera Mod-
ule V2), referred to as RPicams, with a polarising filter (stan-
dard 55 mm circular polariser) attached to one camera. At
present, the two configurations are not compatible to work
concurrently. Recordings using the high-speed camera were
recorded at 1069 frames per second (fps), and recordings us-
ing the RPicams were recorded at 24 fps.
The experimental setup is operated in a cold room which
can achieve a base temperature as low as 50 C and pro-
vided the means of achieving a supercooled environment.
The experimental setup was housed in a Bosch strut and Per-
spex panel frame to prevent the accidental introduction of
frost particles during the experiments. A glass slide was sup-
ported on 3D-printed plastic stilts approximately 10 cm in
height which had a fan attached to dissipate the heat emitted
Atmos. Chem. Phys., 21, 18519–18530, 2021 https://doi.org/10.5194/acp-21-18519-2021
R. L. James et al.: Secondary ice formation during drop–ice collisions 18521
from the polarised light source (LCD liquid-crystal display
monitor). The temperature of the glass slide was monitored
using a K-type thermocouple attached to the glass slide with
aluminium tape. The relative humidity was not measured but
will be below ice saturation, and possibly very small ice
fragments were not observed due to sublimation preventing
growth to visible sizes. The ice particles were prefabricated
by freezing ultrapure water drops (endotoxin-free UP H2O,
Merck) of approximately 6 mm in diameter on a glass slide
coated in a water repellent (Rain-X) using a Peltier cool-
ing system. The typical freezing shape of the ice particle is
shown in Fig. 1. A pipette was modified to allow an ultrapure
water drop (endotoxin-free UP H2O, Merck) at room temper-
ature with a diameter of approximately 5 mm to be placed on
the pipette using a disposable needle (22 gauge, sterile) and
syringe. The modified pipette was held in a 3D-printed tipper
mechanism parallel to the glass slide, and the water drop was
allowed to reach thermal equilibrium with the cold room for
90 s. The supercooled drop was released from the modified
pipette perpendicular to the glass slide and was controlled by
an Arduino and servomotor.
As the drop height and initial supercooled water drop di-
ameter before impact (D) were kept constant at 1.36 m and
5 mm, respectively, the normal impact velocity (V0) for all
experiments was 5.2 m s1. The terminal velocity of a 5 mm
diameter drop is approximately 9 m s1(Gunn and Kinzer,
1949). Initially, the impact velocity may seem unrealistic.
However, the ice particle in these experiments was held sta-
tionary on a glass slide, but in the atmosphere the ice particle
would also be falling. The terminal velocity will depend on
the ice particle shape, but for aggregates of a similar size it is
typically around 1 ms1(Locatelli and Hobbs, 1974). More-
over, turbulence, especially in deep convective clouds, may
also affect the impact velocity (Pinsky and Khain, 1998).
While such large droplets are rare in the atmosphere, the pur-
pose here is to demonstrate that the process is a potential SIP
mechanism. The supercooled water drop and the ice particle
on the glass slide were in thermal equilibrium for all experi-
ments.
The temperature range investigated was between 4 and
12 C. As the temperature of water decreases, the surface
tension (σ) and viscosity (µ) of water increases (Hrubý et al.,
2014; Dehaoui et al., 2015). In fluid dynamics, the We-
ber number, W e =ρDV 2
0 , and Reynolds number, Re =
ρDV0, are used to relate the inertial forces of the fluid
to its interfacial and viscous forces, respectively. In this
case, the fluid is the supercooled water drop, and the di-
ameter of the supercooled water drop, D, refers to the di-
ameter before impact. The inertial force is from the ini-
tial impact velocity of the supercooled water drop, and the
interfacial (surface tension) and viscous forces are proper-
ties of the supercooled water drop. Taking into account the
temperature-dependent values of surface tension and viscos-
ity of the supercooled water between 4 and 12 C, the W e
Figure 1. Schematic diagram of the experimental setup. Compo-
nents labelled (i) were used in the high-speed configuration and (ii)
were used in the RPicam configuration. The setup was operated in
a cold room to achieve a supercooled environment.
and Re number ranges obtained were 1747 W e 1772 and
8781 Re 12 240, respectively.
We conducted 32 experiments using the RPicam configu-
ration during quantification of the freezing fraction of sec-
ondary drops, and the data are given in Table A1.
3 Results
From our high-speed and RPicam recordings we present a
schematic diagram of the formation of secondary drops from
a supercooled water drop impact on an ice particle on a glass
slide in Fig. 2. The W e and R e numbers used were suffi-
ciently large, i.e. W e 2.5 and Re 25, such that inertia
dominated the spreading of the thin film (Roisman, 2009).
Surface tension and viscosity forces were therefore consid-
ered negligible during the spreading phase of the drop (Ro-
isman, 2009), as was the wettability of the surface (Antonini
et al., 2012). Figure 2a depicts the filament-like structures
which were ejected during the spreading phase of the drop
impact. We were unable to track the positions of these sec-
ondary drops or quantify them with our current high-speed
camera or RPicam configurations. As the kinetic energy is
transferred from that of a vertical to horizontal motion at im-
pact, the water drop spread out radially, and instabilities at
the rim were also observed. Figure 2b depicts the retraction
of the drop, which caused the instabilities to “pinch off” or
rupture, followed by a partial rebound. On superhydrophobic
surfaces, rupturing of the instabilities has been attributed to
https://doi.org/10.5194/acp-21-18519-2021 Atmos. Chem. Phys., 21, 18519–18530, 2021
18522 R. L. James et al.: Secondary ice formation during drop–ice collisions
Figure 2. A schematic diagram of a supercooled water drop im-
pact on an ice particle on a glass slide and subsequent secondary
drop formation during (a) the spreading phase and (b) the retraction
phase.
surface tension (Zhang et al., 2020). Our glass slide, coated in
a water repellent, is probably superhydrophobic, and surface
tension is likely the cause of the rupture of the rim instabili-
ties.
3.1 Drop impact: high-speed recordings
We performed control experiments at room temperature
(23 C) and several supercooled temperatures using the high-
speed camera configuration to characterise the water drop
(diameter of 5 mm) impacting the glass slide. Figure 3 shows
the frames from a high-speed recording of (a) a water drop
impact on the glass slide at room temperature and (b) a su-
percooled water drop impact at 5C.
On impact with the glass slide, the water drop deformed
and spread radially outwards as a thin film bordered by a
thicker rim. Instabilities at the rim were observed for both the
room temperature drop and the supercooled drop at 5C.
The supercooled drop shown in Fig. 3b ejected straight
filament-like structures at an angle to the glass surface close
to the impact, and these filament-like structures disintegrated
into secondary drops. This was in contrast to the impact of
the water drop at room temperature where no ejection of
filament-like structures was observed, perhaps due to higher
viscosity and surface tension of water at supercooled temper-
atures. During the retraction phase, some of the rim instabili-
ties pinched off from the thin film in the experiments with the
water drop at room temperature forming secondary drops, in
a process called “receding break-up”. In contrast, no reced-
ing break-up was observed for the supercooled drop.
Figure 4 shows the frames of a supercooled water drop
impacting the side of an ice particle at 5C. Similar to the
supercooled water drop on a glass slide, filament-like struc-
tures, which dissipated into secondary drops, formed at or
close to impact with the ice particle glass slide. Unlike the
impact of a supercooled water drop on a bare glass slide, sec-
ondary drops formed via receding break-up. These secondary
drops were observed around the parts of the rim of the thin
film which contacted the ice particle.
3.2 Determining the freezing fraction of the secondary
drops: RPicams
We performed supercooled water drop impacts on ice parti-
cles between 4 and 12 C. To unambiguously identify if
a secondary drop had frozen, we used a polarising filter with
a polarised light source, exploiting the birefringent proper-
ties of ice. Figure 5 shows selected frames of a supercooled
water drop impact at 4C using the RPicam configuration.
The top row of Fig. 5 shows frames from the camera with no
polarising filter (a) before, (b) at and (c) 10 s after impact.
The number of secondary drops observed is indicated by red
arrows in Fig. 5c. The difference between Fig. 5a and c is
presented in Fig. 5d, clearly indicating the secondary drops
formed. The bottom row shows frames from the camera with
a polarising filter (e) before, (f) at and (g) 10 s after impact.
The frozen secondary drop is indicated by a white arrow in
Fig. 5g. The difference between Fig. 5e and g is presented in
Fig. 5h, clearly indicating the frozen secondary drop formed.
For this particular experiment, five secondary drops
formed, of which one froze, giving a freezing fraction, 8=
0.2. During these experiments, two types of supercooled wa-
ter drop impacts occurred: direct impact on the ice particle
and partial impact on the ice particle. These different impacts
arose due to practical difficulties with consistently impact-
ing the ice particle with the supercooled water drop due to
changes in viscosity of water at different temperatures. For
the experiment shown in Fig. 5, the impact was a side im-
pact towards the top left of the ice particle as indicated in
Fig. 5b. The RPicam configuration only observed the larger
>0.1 mm diameter drops formed during retraction of the thin
film. The smaller secondary drops (<0.1 mm diameter) ob-
served at impact from the high-speed configuration were not
observed using this configuration, as the minimum drop di-
ameter the RPicams could detect was 0.1 mm.
Figure 6a shows the average freezing fraction of secondary
drops formed when a supercooled water drop with a diame-
ter of 5 mm collided with an ice particle, 8, as a function
of temperature. The raw data can be found in Table A1, and
the averaged data of the freezing fraction of secondary drops
are in Table A2. The average number of secondary drops, Ns,
liquid or solid, shown in Fig. 6b as a function of temperature,
reached a maximum at approximately 7.5 C. The averaged
data of the number of secondary drops can be found in Ta-
ble A3.
4 Discussion
We discuss some aspects of the experimental setup that may
affect the occurrence and rate of secondary drop production
and freezing here.
As the ice particles were placed on a flat glass slide, dur-
ing impact, the supercooled water drop spread across the ice
particle and on to the glass slide where the larger >0.1 mm
diameter secondary drops formed. We acknowledge that the
Atmos. Chem. Phys., 21, 18519–18530, 2021 https://doi.org/10.5194/acp-21-18519-2021
R. L. James et al.: Secondary ice formation during drop–ice collisions 18523
Figure 3. Frames from the high-speed camera configuration of a water drop impact on a glass slide when both the water drop and glass slide
are at (a) room temperature (23 C) and (b) 5C. The impact phase (I), spreading phase (S), secondary drop formation/ejection during the
spreading phase (E), retraction phase (R), secondary drop formation due to the receding break-up (B) and partial rebound (PR) of the water
drop are indicated in the frames. Arrows indicate secondary drop formation during the retraction phase of the water drop.
glass slide presents an artificially flat surface compared to at-
mospheric conditions. However, a study by Schremb et al.
(2018) showed that, on an elevated ice surface, the thin film
of a supercooled water drop with a diameter of 4 mm and
similar W e and Re numbers at 14 C was ejected and sub-
sequently ruptured, forming secondary drops. While quan-
tification was not the focus of their study, it was observed that
the rim of the supercooled water drop was largely frozen, and
only some of the secondary drops were observed as ice. The
size of the secondary drops formed in the study by Schremb
et al. (2018) is comparable to our secondary drops despite
the different generation mechanism, and their supplementary
videos indicate that 10 s of secondary drops were formed.
Furthermore, water drops with diameters of 3–4 mm col-
https://doi.org/10.5194/acp-21-18519-2021 Atmos. Chem. Phys., 21, 18519–18530, 2021
18524 R. L. James et al.: Secondary ice formation during drop–ice collisions
Figure 4. Frames from the high-speed camera configuration of a supercooled water drop impact on an ice particle when both the drop and ice
particle are at 5C. The impact phase (I), spreading phase (S), secondary drop formation/ejection during the spreading phase (E), retraction
phase (R), secondary drop formation due to the receding break-up (B) and partial rebound (PR) of the water drop are indicated in the frames.
Arrows indicate secondary drop formation during the retraction phase of the supercooled water drop.
Figure 5. Selected frames from the impact of a supercooled water drop on an ice particle at 4C using the RPicam configuration. Frames
(a)(c) before, at and 10 s after impact using the camera with no polarising filter. Red arrows in (c) indicate the number of secondary
drops formed. Frame (d) shows the difference between (a) and (c). Frames (e)(g) before, at and 10 s after impact using the camera with a
polarising filter. The white arrow in (h) indicates the frozen secondary drop. Frame (h) shows the difference between (e) and (g).
liding with a steel disk of 4 mm in diameter (Rozhkov
et al., 2002) and water drops with diameters of 6 mm col-
liding with an iron cylinder of the same diameter (Viller-
maux and Bossa, 2011) produced 100 s of secondary drops.
Clearly, secondary drops still form, emitted from the rim of
the thin film during impact, when there is no supporting flat
surface, such as the glass slide used in this study. However,
there is much uncertainty about the number of secondary
drops formed.
In addition, the ice particle in our experiments is in a fixed
position on the glass slide, whereas, in the atmosphere, the
ice particle is in free fall. When the faster-moving super-
Atmos. Chem. Phys., 21, 18519–18530, 2021 https://doi.org/10.5194/acp-21-18519-2021
R. L. James et al.: Secondary ice formation during drop–ice collisions 18525
Figure 6. (a) The average freezing fraction of the secondary drops
(8, black triangles) and (b) the average number of secondary drops
(Ns, blue circles) as a function of temperature. Average data in-
cluded both direct and partial collisions. The error bars represent
the standard error in the freezing fraction or secondary drop num-
ber for the temperature intervals which are listed in Tables A2 and
A3.
cooled water drop collides with the ice particle, the ice parti-
cle will move in response to the collision, likely affecting the
formation of the secondary drops and their subsequent freez-
ing. However, currently, it is difficult to ascertain how this
will influence secondary drop formation and freezing with-
out further investigations into the mechanisms of secondary
drop formation on an elevated ice particle.
Another factor that will influence the generation of sec-
ondary drops is the ice particle shape. Our ice particles have
a pointed tip, as shown in Fig. 2, which is a typical shape
formed when a liquid water drop is frozen on a cold sub-
strate (Snoeijer and Brunet, 2012) but not representative of
atmospheric ice particles. According to Phillips et al. (2018),
who refer to this SIP mechanism as mode 2, for it to occur,
the supercooled water drops must have a diameter larger than
150 µm, and the ice particle must be more massive still. In the
atmosphere, ice particles which are larger than 150 µm are
typically irregular in shape (Korolev and Sussman, 2000). A
study by Zhang et al. (2020) shows that at room temperature,
water drop impact on curved surfaces induce additional frag-
mentation mechanisms compared to flat surfaces. Therefore,
we expect the irregular shape of an ice particle to introduce
additional fragmentation mechanisms of the supercooled wa-
ter drop, which may enhance secondary drop formation.
We observed a decrease in the number of secondary drops
formed during receding break-up as the temperature de-
creased below 8C. Figure 7 shows the frames after a su-
percooled water drop impact with an ice particle for the ex-
periments between 11 and 12 C, which was the range
where the smallest number of secondary drops formed. At
these temperatures, the supercooled water drop froze either
during the spreading phase or in the early stages of the re-
traction phase. The growth velocity of ice in supercooled wa-
ter increases with decreasing temperature (e.g. at 2C it is
around 0.2 cm s1, whereas at 10 C it is around 5 cm s1;
Pruppacher, 1967), which may explain why a decrease in sec-
ondary drops was observed. We believe the decrease in sec-
ondary drop formation at temperatures below 8C may be
due to the artificially flat geometry presented by the glass
slide and to the large size of the incident drop, both fac-
tors which prolonged the interaction time between the su-
percooled water drop and ice. For example, the supplemen-
tary videos from Schremb et al. (2018) showed 10 s of sec-
ondary drops forming at 14 C after impact on an elevated
ice target, more than we observed at our lowest temperature
of 12 C.
Whilst the freezing mechanism of the secondary drops was
not specifically studied in this work, we consider the follow-
ing mechanisms. The freezing of supercooled water drops
occurs in two stages. The first stage is characterised by the
formation of ice dendrites throughout the supercooled water
drop. The latent heat from the formation of the ice dendrites
is released during this stage, warming the temperature of the
supercooled water drop to 0C. The second stage is char-
acterised by the freezing of the remaining supercooled water
drop and is controlled by the loss of latent heat due to the sur-
roundings of the supercooled water drop. Stage 1 of freezing
is fast, and the time taken for this stage to complete (ti) can be
estimated from the following equation (Macklin and Payne,
1967):
tiδR
G,(1)
where δRis the thickness of the layer of supercooled water
on the ice particle and Gis the growth velocity of ice, which
is temperature dependent.
From Fig. 4, we can estimate that the rim of the super-
cooled water drop, which is also the thickest part of the su-
percooled water drop, is approximately 0.78 mm. Taking this
value for δRand given that the growth velocity of ice at
5C is approximately 1 cm s1(Pruppacher, 1967), then
ti0.078 s. Figure 4 shows the timescale for the retraction
phase is of the order of 0.1 s. It is plausible that the initial ice
dendrites can propagate through the supercooled water drop
and that water containing these ice dendrites may then break
off during the retraction phase and initiate freezing. The sec-
ond phase of freezing will take longer, but as long as the drop
contains ice dendrites it will eventually freeze. This explana-
tion is also proposed by Schremb et al. (2018) and Phillips
https://doi.org/10.5194/acp-21-18519-2021 Atmos. Chem. Phys., 21, 18519–18530, 2021
18526 R. L. James et al.: Secondary ice formation during drop–ice collisions
Figure 7. Frames from the RPicam configuration approximately 10 s after a supercooled water drop impact for experiments at T > 11 C.
The top panel shows frames from the RPicam with no polarising filter, and the bottom panel shows frames from the RPicam with a polarising
filter.
et al. (2018), who suggest that seeding ice crystals are trans-
ported during the initial spreading phase. Alternative freez-
ing mechanisms include the formation of a thin, unobserved
film of liquid water present on the glass slide after the retrac-
tion phase. The contact between the thin film of water and
the ice particle could induce freezing in the thin film, which
could then trigger freezing in the seemingly detached sec-
ondary drop. Mechanical agitation or shock may also play
a role in the freezing of the secondary drops (Alkezweeny,
1969; Czys, 1989). Regardless of the freezing mechanism,
the glass slide will likely have some influence, and it will be
pertinent to remove this in future investigations.
As a proof-of-concept investigation, we studied super-
cooled water drops with diameters of 5 mm and ice parti-
cles with diameters of 6 mm, as larger sizes of supercooled
water drops were easier to work with experimentally. While
these sizes are not necessarily representative of cloud con-
ditions, theoretically, this new SIP mechanism should occur
where supercooled water drop diameters are >150 µm and
the ice particles are more massive still. Supercooled water
drops and ice particles are present within a variety of dif-
ferent clouds. For example, Hobbs and Rangno (1990) pre-
sented aircraft observations in small polar–maritime cumuli
that displayed ice enhancement with cloud bases too cold
for large cloud droplets (>24 µm) between 3 and 8C
as required for rime splintering. Their discussion highlighted
that ice enhancement proceeded in two stages. The first stage
consisted of the formation of frozen drops, <400 µm diam-
eter, and small graupel particles, <1 mm diameter. The sec-
ond stage was characterised by the appearance of high con-
centrations of vapour-grown ice crystals in the upper regions
of the cloud. A key finding of this series of papers was that
high concentrations of small ice particles appeared simulta-
neously with frozen drizzle drops. Furthermore, Rangno and
Hobbs (2001) showed that large supercooled drops were of-
ten a requirement for ice enhancement in moderately cooled
Arctic stratiform clouds, and ice enhancement was often co-
incident with observations of large supercooled raindrops.
Supercooled drizzle drops and raindrops are common in con-
vective clouds (e.g. Crawford et al., 2012; Taylor et al.,
2016), as are large ice particles. Hence, the broad continuum
of drizzle and raindrop sizes, where the larger drops freeze
first, followed by accretion of the smaller unfrozen drops in-
dicates that collisions of supercooled water drops with ice
particles that are more massive may be of importance in a
wide range of clouds.
5 Conclusions and future work
In this study, we confirmed that during collisions of super-
cooled water drops with ice particles, frozen secondary drops
formed over the temperature range between 4 and 12 C.
Our main findings are the following:
1. Approximately 5 to 10 secondary drops are formed by
receding break-up during the retraction phase of a su-
percooled water drop (D=5 mm) after collision with
an ice particle (D=6 mm) placed on a glass slide.
2. An average of 30 % of these secondary drops formed
froze between 4 and 12 C.
3. Experiments with a high-speed camera highlighted that
secondary drops formed as a jet of smaller droplets
produced separately from the receding break-up of the
drop. No quantification of the freezing fraction of the
secondary drops can currently be made.
One of the main experimental challenges of this work was
dropping the supercooled water drop consistently onto the
ice particle, which limited the number of experiments we
could perform. As shown in Table A1, the majority of the
successful impacts were classified as partial hits despite the
intention for them to be direct hits. While partial hits are ex-
pected in clouds, as well as direct hits, we also conducted
Atmos. Chem. Phys., 21, 18519–18530, 2021 https://doi.org/10.5194/acp-21-18519-2021
R. L. James et al.: Secondary ice formation during drop–ice collisions 18527
many experiments where the supercooled water drop missed
the ice particle. One method of achieving better control of
the supercooled water drop impact could be via growth and
supercooling of a water drop at the end of a needle similar
to the system shown in Schremb et al. (2018). Compared to
our current mechanism, which involved tilting a pipette to
allow the supercooled water drop to roll off, the supercooled
water drop would remain fixed to a certain point before de-
taching under gravity, making it easier to drop consistently
in the same position.
Another experimental challenge we would like to address
is quantifying the secondary drops formed during the spread-
ing phase of the supercooled water drop during impact.
Thoroddsen et al. (2012) quantified secondary drops ejected
with velocities of up to 100 m s1using an ultra-high-speed
camera capable of recording at 1 000 000 fps, and we could
use a similar setup. We could then exploit the birefringent
properties of ice to determine whether these ejected sec-
ondary drops froze.
The number of secondary drops per collision is sensitive to
geometry and the material of collision, even for drops of the
same size. We quantify about 10 secondary drops per col-
lision. Schremb et al. (2018) observe of the order of tens
of drops per collision for impacts on elevated ice surface.
Finally, Rozhkov et al. (2002) observe 100 s for drop im-
pacts on steel disks at room temperature, as do Villermaux
and Bossa (2011) for drop impacts on iron cylinders at room
temperatures. Consequently, after addressing the above chal-
lenges and elevating the ice particle off the glass surface,
which may be achieved simply by fixing the ice particle on a
wire, further work is needed to investigate, more systemati-
cally, this new SIP mechanism over a range of experimental
parameters, not limited to the following: supercooled drop
sizes, size ratios of supercooled water drops to ice particles,
ice particle shapes, temperatures, drop height (and hence
impact velocity), airflow, relative humidity conditions and
chemical compositions of the supercooled water drop.
Appendix A
Table A1. Total number of secondary drops and the number of
frozen secondary drops for each experiment.
Temperature Total number of Frozen secondary
(C) secondary drops drops
Direct 4.2 14 1
4.2 0 0
5.3 7 6
5.5 10 2
7.8 12 2
9.9 7 0
Partial 3.8 16 5
4.0 5 1
4.0 8 5
4.3 9 6
5.6 5 0
5.6 5 1
5.8 9 5
6.0 4 1
6.0 8 2
6.1 2 1
6.1 12 3
7.7 17 7
8.0 5 0
8.0 11 7
8.1 8 1
8.5 16 0
9.4 0 0
9.4 21 6
9.8 11 4
10.0 2 2
10.1 10 6
11.3 0 0
11.5 4 1
11.8 4 1
11.9 0 0
11.9 0 0
https://doi.org/10.5194/acp-21-18519-2021 Atmos. Chem. Phys., 21, 18519–18530, 2021
18528 R. L. James et al.: Secondary ice formation during drop–ice collisions
Table A2. The mean (8) and standard deviation (SD) of the fraction
of frozen secondary drops within a specified temperature interval
(Tinterval) along with the number of experiments (n) within the T
interval, the average degree of supercooling within the temperature
interval (T) and the error in the sample mean (σ8).
Tinterval (C) T(C) 8SD n σ8
3.8 4.3 4.1 0.38 0.26 5 0.12
5.3 5.8 5.6 0.36 0.34 5 0.15
6.0 6.1 6.1 0.44 0.38 4 0.19
7.7 7.8 7.8 0.29 0.17 2 0.12
8.0 8.5 8.2 0.19 0.30 4 0.15
9.4 9.9 9.7 0.22 0.19 3 0.11
10.0 10.1 10.1 0.80 0.28 2 0.20
11.3 11.9 11.7 0.25 0.00 2 0.00
Table A3. The mean (Ns) and standard deviation (SD) of the num-
ber of secondary drops within a specified temperature interval (T
interval) along with the number of experiments (n) within the Tin-
terval, the average degree of supercooling within the temperature
interval (T) and the error in the sample mean (σNs).
Tinterval (C) T(C) NsSD n σNs
3.8 4.3 4.1 8.7 5.9 6 2.4
5.3 5.8 5.6 7.2 2.3 5 1.0
6.0 6.1 6.1 6.5 4.4 4 2.2
7.7 7.8 7.8 14.5 3.5 2 2.5
8.0 8.5 8.2 10.0 4.7 4 2.3
9.4 9.9 9.6 9.8 8.8 4 4.4
10.0 10.1 10.1 6.0 5.7 2 4.0
11.3 11.9 11.7 1.6 2.2 5 1.0
Data availability. All datasets are provided in Appendix A.
Video supplement. All video recordings from the high-
speed configuration and RPicam configuration are deposited
in Figshare, a FAIR-aligned (findable, accessible, interoper-
able and re-usable) data repository, and can be accessed at
https://doi.org/10.48420/c.5476557 (James et al., 2021).
Author contributions. VTJP and PJC conceived the original
study. RLJ and PJC designed the new experimental setup with ad-
vice from VTJP. RLJ and PJC performed the experiments. RLJ anal-
ysed the data and wrote the paper. VTJP and PJC provided com-
ments on the paper.
Competing interests. The contact author has declared that nei-
ther they nor their co-authors have any competing interests.
Disclaimer. Publisher’s note: Copernicus Publications remains
neutral with regard to jurisdictional claims in published maps and
institutional affiliations.
Acknowledgements. We thank Sylvia Sullivan; the two anony-
mous reviewers; and the editor, Timothy Garrett, whose comments
helped significantly improve the quality of this paper.
Financial support. This research has been supported by the U.S.
Department of Energy (grant no. DE-SC0018932) and the Natural
Environment Research Council (grant no. NE/T001496/1).
Review statement. This paper was edited by Timothy Garrett and
reviewed by Sylvia Sullivan and two anonymous referees.
References
Alkezweeny, A. J.: Freezing of Supercooled Wa-
ter Droplets due to Collision, J. Appl. Meteo-
rol. Clim., 8, 994–995, https://doi.org/10.1175/1520-
0450(1969)008<0994:FOSWDD>2.0.CO;2, 1969.
Antonini, C., Amirfazli, A., and Marengo, M.: Drop impact and
wettability: From hydrophilic to superhydrophobic surfaces,
Phys. Fluids, 24, 102104, https://doi.org/10.1063/1.4757122,
2012.
Blyth, A. M. and Latham, J.: Development of ice and precipitation
in new mexican summertime cumulus, Q. J. Roy. Meteor. Soc.,
119, 91–120, https://doi.org/10.1002/qj.49711950905, 1993.
Blyth, A. M. and Latham, J.: A multi-thermal model of cumulus
glaciation via the Hallett–Mossop process, Q. J. Roy. Meteor.
Soc., 123, 1185–1198, https://doi.org/10.1002/qj.49712354104,
1997.
Changnon, S. A.: Characteristics of Ice Storms in the United States,
J. Appl. Meteorol., 42, 630–639, https://doi.org/10.1175/1520-
0450(2003)042<0630:COISIT>2.0.CO;2, 2003.
Crawford, I., Bower, K. N., Choularton, T. W., Dearden, C., Crosier,
J., Westbrook, C., Capes, G., Coe, H., Connolly, P. J., Dorsey,
J. R., Gallagher, M. W., Williams, P., Trembath, J., Cui, Z.,
and Blyth, A.: Ice formation and development in aged, win-
tertime cumulus over the UK: observations and modelling, At-
mos. Chem. Phys., 12, 4963–4985, https://doi.org/10.5194/acp-
12-4963-2012, 2012.
Crosier, J., Bower, K. N., Choularton, T. W., Westbrook, C. D., Con-
nolly, P. J., Cui, Z. Q., Crawford, I. P., Capes, G. L., Coe, H.,
Dorsey, J. R., Williams, P. I., Illingworth, A. J., Gallagher, M. W.,
and Blyth, A. M.: Observations of ice multiplication in a weakly
convective cell embedded in supercooled mid-level stratus, At-
mos. Chem. Phys., 11, 257–273, https://doi.org/10.5194/acp-11-
257-2011, 2011.
Czys, R. R.: Ice Initiation by Collision-Freezing in
Warm-Based Cumuli, J. Appl. Meteorol. Clim.,
28, 1098–1104, https://doi.org/10.1175/1520-
0450(1989)028<1098:IIBCFI>2.0.CO;2, 1989.
Dehaoui, A., Issenmann, B., and Caupin, F.: Viscosity of
deeply supercooled water and its coupling to molecular
Atmos. Chem. Phys., 21, 18519–18530, 2021 https://doi.org/10.5194/acp-21-18519-2021
R. L. James et al.: Secondary ice formation during drop–ice collisions 18529
diffusion, P. Natl. Acad. Sci. USA, 112, 12020–12025,
https://doi.org/10.1073/pnas.1508996112, 2015.
Elsom, D.: Deaths and injuries caused by lightning in the United
Kingdom: analyses of two databases, Atmos. Res., 56, 325–334,
https://doi.org/10.1016/S0169-8095(00)00083-1, 2001.
Field, P. R. and Heymsfield, A. J.: Importance of snow to
global precipitation, Geophys. Res. Lett., 42, 9512–9520,
https://doi.org/10.1002/2015GL065497, 2015.
Field, P. R., Lawson, R. P., Brown, P. R. A., Lloyd, G., West-
brook, C., Moisseev, D., Miltenberger, A., Nenes, A., Blyth, A.,
Choularton, T., Connolly, P., Buehl, J., Crosier, J., Cui, Z., Dear-
den, C., DeMott, P., Flossmann, A., Heymsfield, A., Huang, Y.,
Kalesse, H., Kanji, Z. A., Korolev, A., Kirchgaessner, A., Lasher-
Trapp, S., Leisner, T., McFarquhar, G., Phillips, V., Stith, J., and
Sullivan, S.: Secondary Ice Production: Current State of the Sci-
ence and Recommendations for the Future, Meteorol. Monogr.,
58, 7.1–7.20, https://doi.org/10.1175/AMSMONOGRAPHS-D-
16-0014.1, 2017.
Gunn, R. and Kinzer, G. D.: The terminal veloc-
ity of fall for water droplets in stagnant air, J. At-
mos. Sci., 6, 243–248, https://doi.org/10.1175/1520-
0469(1949)006<0243:TTVOFF>2.0.CO;2, 1949.
Hallett, J. and Mossop, S. C.: Production of secondary ice
particles during the riming process, Nature, 249, 26–28,
https://doi.org/10.1038/249026a0, 1974.
Harris-Hobbs, R. L. and Cooper, W. A.: Field Evidence Support-
ing Quantitative Predictions of Secondary Ice Production Rates,
J. Atmos. Sci., 44, 1071–1082, https://doi.org/10.1175/1520-
0469(1987)044<1071:FESQPO>2.0.CO;2, 1987.
Hobbs, P. V. and Rangno, A. L.: Ice parti-
cle concentrations in clouds, J. Atmos. Sci.,
42, 2523–2549, https://doi.org/10.1175/1520-
0469(1985)042<2523:IPCIC>2.0.CO;2, 1985.
Hobbs, P. V. and Rangno, A. L.: Rapid development of high ice par-
ticle concentrations in small polar maritime cumuliform clouds,
J. Atmos. Sci., 47, 2710–2722, https://doi.org/10.1175/1520-
0469(1990)047<2710:RDOHIP>2.0.CO;2, 1990.
Hrubý, J., Vinš, V., Mareš, R., Hykl, J., and Kalová, J.:
Surface Tension of Supercooled Water: No Inflection
Point down to 25 C, J. Phys. Chem. Lett., 5, 425–428,
https://doi.org/10.1021/jz402571a, 2014.
Huang, Y., Blyth, A. M., Brown, P. R. A., Choularton, T. W.,
and Cui, Z.: Factors controlling secondary ice production in
cumulus clouds, Q. J. Roy. Meteor. Soc., 143, 1021–1031,
https://doi.org/10.1002/qj.2987, 2017.
James, R., Phillips, V. T. J., and Connolly, P. J.: Secondary ice pro-
duction during the break-up of freezing water drops on impact
with ice particles, University of Manchester [data set], Collec-
tion, https://doi.org/10.48420/c.5476557.v1, 2021.
Kanji, Z. A., Ladino, L. A., Wex, H., Boose, Y., Burkert-
Kohn, M., Cziczo, D. J., and Krämer, M.: Overview of
Ice Nucleating Particles, Meteorol. Monogr., 58, 1.1–1.33,
https://doi.org/10.1175/AMSMONOGRAPHS-D-16-0006.1,
2017.
Korolev, A. and Leisner, T.: Review of experimental studies of sec-
ondary ice production, Atmos. Chem. Phys., 20, 11767–11797,
https://doi.org/10.5194/acp-20-11767-2020, 2020.
Korolev, A. and Sussman, B.: A Technique for Habit
Classification of Cloud Particles, J. Atmos. Ocean.
Tech., 17, 1048–1057, https://doi.org/10.1175/1520-
0426(2000)017<1048:ATFHCO>2.0.CO;2, 2000.
Ladino, L. A., Korolev, A., Heckman, I., Wolde, M., Fridlind,
A. M., and Ackerman, A. S.: On the role of ice-nucleating
aerosol in the formation of ice particles in tropical mesoscale
convective systems, Geophys. Res. Lett., 44, 1574–1582,
https://doi.org/10.1002/2016GL072455, 2017.
Lasher-Trapp, S., Leon, D. C., DeMott, P. J., Villanueva-Birriel,
C. M., Johnson, A. V., Moser, D. H., Tully, C. S., and
Wu, W.: A Multisensor Investigation of Rime Splintering in
Tropical Maritime Cumuli, J. Atmos. Sci., 73, 2547–2564,
https://doi.org/10.1175/JAS-D-15-0285.1, 2016.
Latham, J. and Warwicker, R.: Charge transfer accom-
panying the splashing of supercooled raindrops on
hailstones, Q. J. Roy. Meteor. Soc., 106, 559–568,
https://doi.org/10.1002/qj.49710644912, 1980.
Lloyd, G., Choularton, T. W., Bower, K. N., Gallagher, M. W.,
Connolly, P. J., Flynn, M., Farrington, R., Crosier, J., Sch-
lenczek, O., Fugal, J., and Henneberger, J.: The origins of
ice crystals measured in mixed-phase clouds at the high-
alpine site Jungfraujoch, Atmos. Chem. Phys., 15, 12953–12969,
https://doi.org/10.5194/acp-15-12953-2015, 2015.
Locatelli, J. D. and Hobbs, P. V.: Fall speeds and masses of
solid precipitation particles, J. Geophys. Res., 79, 2185–2197,
https://doi.org/10.1029/JC079i015p02185, 1974.
Macklin, W. C. and Payne, G. S.: A theoretical study of the
ice accretion process, Q. J. Roy. Meteor. Soc., 93, 195–213,
https://doi.org/10.1002/qj.49709339606, 1967.
Mossop, S. C.: Some Factors Governing Ice Parti-
cle Multiplication in Cumulus Clouds, J. Atmos.
Sci., 35, 2033–2037, https://doi.org/10.1175/1520-
0469(1978)035<2033:SFGIPM>2.0.CO;2, 1978.
Mossop, S. C. and Hallett, J.: Ice Crystal Concentration in Cumulus
Clouds: Influence of the Drop Spectrum, Science, 186, 632–634,
https://doi.org/10.1126/science.186.4164.632, 1974.
Phillips, V. T., Patade, S., Gutierrez, J., and Bansemer, A.: Sec-
ondary Ice Production by Fragmentation of Freezing Drops:
Formulation and Theory, J. Atmos. Sci., 76, 3031–3070,
https://doi.org/10.1175/JAS-D-17-0190.1, 2018.
Phillips, V. T. J., Blyth, A. M., Brown, P. R. A., Choularton,
T. W., and Latham, J.: The glaciation of a cumulus cloud
over New Mexico, Q. J. Roy. Meteor. Soc., 127, 1513–1534,
https://doi.org/10.1002/qj.49712757503, 2001.
Phillips, V. T. J., Andronache, C., Sherwood, S. C., Bansemer, A.,
Conant, W. C., Demott, P. J., Flagan, R. C., Heymsfield, A.,
Jonsson, H., Poellot, M., Rissman, T. A., Seinfeld, J. H., Van-
reken, T., Varutbangkul, V., and Wilson, J. C.: Anvil glacia-
tion in a deep cumulus updraught over Florida simulated with
the Explicit Microphysics Model. I: Impact of various nu-
cleation processes, Q. J. Roy. Meteor. Soc., 131, 2019–2046,
https://doi.org/10.1256/qj.04.85, 2005.
Pinsky, M. and Khain, A.: Some effects of cloud turbulence on
water–ice and ice–ice collisions, Atmos. Res., 47–48, 69–86,
https://doi.org/10.1016/S0169-8095(98)00041-6, 1998.
Pruppacher, H. R.: On the growth of ice crystals in supercooled wa-
ter and aqueous solution drops, Pure Appl. Geophys., 68, 186–
195, https://doi.org/10.1007/BF00874894, 1967.
ˇ
cik, T., Castellano, C., Groenemeijer, P., Kühne, T., Rädler,
A. T., Antonescu, B., and Faust, E.: Large Hail Incidence
https://doi.org/10.5194/acp-21-18519-2021 Atmos. Chem. Phys., 21, 18519–18530, 2021
18530 R. L. James et al.: Secondary ice formation during drop–ice collisions
and Its Economic and Societal Impacts across Europe, Mon.
Weather Rev., 147, 3901–3916, https://doi.org/10.1175/MWR-
D-19-0204.1, 2019.
Rangno, A. L. and Hobbs, P. V.: Ice particles in stratiform clouds
in the Arctic and possible mechanisms for the production of
high ice concentrations, J. Geophys. Res., 106, 15065–15075,
https://doi.org/10.1029/2000JD900286, 2001.
Roisman, I. V.: Inertia dominated drop collisions. II. An
analytical solution of the Navier–Stokes equations for
a spreading viscous film, Phys. Fluids, 21, 052104,
https://doi.org/10.1063/1.3129283, 2009.
Rozhkov, A., Prunet-Foch, B., and Vignes-Adler, M.: Impact of
water drops on small targets, Phys. Fluids, 14, 3485–3501,
https://doi.org/10.1063/1.1502663, 2002.
Schremb, M., Roisman, I. V., and Tropea, C.: Normal impact
of supercooled water drops onto a smooth ice surface: ex-
periments and modelling, J. Fluid Mech., 835, 1087–1107,
https://doi.org/10.1017/jfm.2017.797, 2018.
Snoeijer, J. H. and Brunet, P.: Pointy ice-drops: How water
freezes into a singular shape, Am. J. Phys., 80, 764–771,
https://doi.org/10.1119/1.4726201, 2012.
Taylor, J. W., Choularton, T. W., Blyth, A. M., Liu, Z., Bower, K.
N., Crosier, J., Gallagher, M. W., Williams, P. I., Dorsey, J. R.,
Flynn, M. J., Bennett, L. J., Huang, Y., French, J., Korolev, A.,
and Brown, P. R. A.: Observations of cloud microphysics and
ice formation during COPE, Atmos. Chem. Phys., 16, 799–826,
https://doi.org/10.5194/acp-16-799-2016, 2016.
Thoroddsen, S. T., Takehara, K., and Etoh, T. G.: Micro-
splashing by drop impacts, J. Fluid Mech., 706, 560–570,
https://doi.org/10.1017/jfm.2012.281, 2012.
Villermaux, E. and Bossa, B.: Drop fragmenta-
tion on impact, J. Fluid Mech., 668, 412–435,
https://doi.org/10.1017/S002211201000474X, 2011.
Westbrook, C. D. and Illingworth, A. J.: The formation of ice in a
long-lived supercooled layer cloud, Q. J. Roy. Meteor. Soc., 139,
2209–2221, https://doi.org/10.1002/qj.2096, 2013.
Zhang, H., Zhang, X., Yi, X., He, F., Niu, F., and Hao, P.:
Asymmetric splash and breakup of drops impacting on cylin-
drical superhydrophobic surfaces, Phys. Fluids, 32, 122108,
https://doi.org/10.1063/5.0032910, 2020.
Atmos. Chem. Phys., 21, 18519–18530, 2021 https://doi.org/10.5194/acp-21-18519-2021
... Ice multiplication, also known as SIP, can be an important source of ice particles in MPCs, as it can rapidly increase the pre-existing ice crystal number concentrations (ICNCs) through a number of collisional processes (Field et al., 2017;Korolev & Leisner, 2020). The most frequently acknowledged SIP mechanisms are the Hallett-Mossop rime splintering (HM; Choularton et al., 1980;Hallett & Mossop, 1974), ice-ice collisional break-up (BR; Takahashi et al., 1995;Vardiman, 1978) and droplet-shattering (DS; Choularton et al., 1980;James et al., 2021;Kleinheins et al., 2021). Observations of MPCs in the Arctic region have shown high ICNCs that greatly surpass ice nucleating particle (INP) concentrations (Wex et al., 2019), particularly at temperatures above 25°C (Luke et al., 2021;Pasquier et al., 2022;Rangno & Hobbs, 2001;Schwarzenboeck et al., 2009). ...
... The shattering probability depends on raindrop size, being 0 for sizes smaller than 50 μm, 1 for sizes larger than 60 μm, and varying for sizes in between. The second mode, or "mode 2," involves collisions between raindrops and larger ice particles such as snow or graupel (James et al., 2021), producing only tiny ice fragments that are passed to the cloud ice category. Big fragments are treated as either graupel, snow, or frozen drop depending on the type of collision that initiated the raindrop freezing process. ...
... In mode 2, the collision with more massive ice particles disrupts the spherical symmetry of the raindrop, generating only tiny fragments upon freezing. Although there is only one dedicated laboratory study on this freezing mode (James et al., 2021), Phillips et al. (2018) proposed a theoretical, energy-based formulation to represent the number of tiny splinters per drop accreted: ...
Article
Full-text available
Accurately representing mixed‐phase clouds (MPCs) in global climate models (GCMs) is critical for capturing climate sensitivity and Arctic amplification. Secondary ice production (SIP), can significantly increase ice crystal number concentration (ICNC) in MPCs, affecting cloud properties and processes. Here, we introduce a machine‐learning (ML) approach, called Random Forest SIP (RaFSIP), to parameterize SIP in stratiform MPCs. RaFSIP is trained on 16 grid points with 10‐km horizontal spacing derived from a 2‐year simulation with the Weather Research and Forecasting (WRF) model, including explicit SIP microphysics. Designed for a temperature range of 0 to −25°C, RaFSIP simplifies the description of rime splintering, ice‐ice collisional break‐up, and droplet‐shattering using only a limited set of inputs. RaFSIP was evaluated offline before being integrated into WRF, demonstrating its stable online performance in a 1‐year simulation keeping the same model setup as during training. Even when coupled with the 50‐km grid spacing domain of WRF, RaFSIP reproduces ICNC predictions within a factor of 3 when compared to simulations with explicit SIP microphysics. The coupled WRF‐RaFSIP scheme replicates regions of enhanced SIP and accurately maps ICNCs and liquid water content, particularly at temperatures above −10°C. Uncertainties in RaFSIP minimally impact surface cloud radiative forcing in the Arctic, resulting in radiative biases under 3 Wm⁻² compared to simulations with detailed microphysics. Although the performance of RaFSIP in convective clouds remains untested, its adaptable nature allows for data set augmentation to address this aspect. This framework opens possibilities for GCM simplification and process description through physics‐guided ML algorithms.
... The importance of SIP has been widely acknowledged in laboratory [23][24][25][26] , eld [27][28][29] , remote sensing [30][31][32][33] , and modeling studies [34][35][36] worldwide 37 . The most commonly invoked SIP processes include the Hallett-Mossop (HM) or rime-splintering process 38,39 , ice-ice collisional break-up (BR) 40,41 , and dropletshattering during freezing (DS) 42,43 . ...
... The former is set to unity for temperatures below -6°C and zero for temperatures above -3°C, while the latter depends on the size of the raindrop, being 0 for sizes smaller than 50 μm, 1 for sizes larger than 60 μm. The second mode, involves collisions between raindrops and larger ice particles such as snow or graupel 25 . These collisions produce tiny ice fragments, which are introduced as cloud ice in the number conservation equations. ...
Preprint
Full-text available
Recent years have shown that secondary ice production (SIP) is ubiquitous, affecting all clouds from polar to tropical regions. SIP is not described well in models and for this may vastly underpredict ice crystal number concentrations in warm mixed-phase clouds. Through a synergy of modeling, remote sensing and in-situ measurements carried out in an orographic environment during the Cloud-AerosoL InteractionS in the Helmos background TropOsphere (CALISHTO) campaign, we show that SIP can have a profound impact on the vertical distribution of hydrometeors and precipitation, especially in seeder-feeder configurations which are encountered in multi-layered cloud systems. The mesoscale model simulations coupled with a radar simulator strongly support a unique signature that is characteristic of SIP; because of this, our study opens the possibility of using the vast global archive of cloud radar data for systematically inferring SIP signatures and frequency of occurrence.
... The former is set to unity for temperatures below −6°C and zero for temperatures above −3°C, while the latter depends on the size of the raindrop, being 0 for sizes smaller than 50 μm, 1 for sizes larger than 60 μm. The second mode, involves collisions between raindrops and larger ice particles such as snow or graupel 92 . These collisions produce tiny ice fragments, which are introduced as cloud ice in the number conservation equations. ...
Article
Full-text available
Recent years have shown that secondary ice production (SIP) is ubiquitous, affecting all clouds from polar to tropical regions. SIP is not described well in models and may explain biases in warm mixed-phase cloud ice content and structure. Through modeling constrained by in-situ observations and its synergy with radar we show that SIP in orographic clouds exert a profound impact on the vertical distribution of hydrometeors and precipitation, especially in seeder-feeder cloud configurations. The mesoscale model simulations coupled with a radar simulator strongly support that enhanced aggregation and SIP through ice-ice collisions contribute to observed spectral bimodalities, skewing the Doppler spectra toward the slower-falling side at temperatures within the dendritic growth layer, ranging from −20 °C to −10 °C. This unique signature provides an opportunity to infer long-term SIP occurrences from the global cloud radar data archive, particularly for this underexplored temperature regime.
... C w and L f are the specific heat capacity of water and the specific latent heat of freezing, respectively. DE crit = 0.2, and is 0.3 according to James et al. (2021). All ice fragments are assumed to be tiny in this mode. ...
Article
Full-text available
Ice microphysics controls cloud electrification in thunderstorms, and the various secondary ice production (SIP) processes are vital in generating high ice concentrations. However, the role of SIP in cold-season thunderstorms is not well understood. In this study, the impacts of SIP on the electrification in a thunderstorm that occurred in late November are investigated using model simulations. The parameterizations of four SIP processes are implemented in the model, including the rime splintering, ice–ice collisional breakup, shattering of freezing drops, and sublimational breakup of ice. In addition, a noninductive charging parameterization and an inductive charging parameterization, as well as a bulk discharging model, are coupled with the spectral bin microphysics scheme. The macroscopic characteristics and the temporal evolution of this thunderstorm are well modeled. The radar reflectivity and flash rate obtained by adding four SIP processes are more consistent with the observations than those without SIP. Among the four SIP processes, the rime splintering has the strongest impact on the storm. The graupel and snow concentrations are enhanced while their sizes are suppressed due to the SIP. The changes in the ice microphysics result in substantial changes in the charge structure. The total charge density changes from an inverted tripole structure to a dipole structure (tripole structure at some locations) after four SIP processes are considered in the model, mainly due to the enhanced collision between graupel and ice. These changes lead to an enhancement of the vertical electric field, especially in the mature stage, which explains the improved modeling of flash rate. The results highlight that cold-season cloud electrification is very sensitive to the SIP processes.
... The latter is crucial for detailed studies of the dependence of FFD rates on environmental conditions and the development of physicallybased parameterizations that can be used in numerical simulations. The simulations examined in this study support the hypothesis that FFD SIP alone is sufficient to explain the large amounts 1185 of ice generated without involving other SIP mechanisms (Korolev and Leisner, 2020;James et al., 2021). That said, the authors acknowledge that the FFD mechanism is not well established in the laboratory, and the contribution of other mechanisms also remains uncertain and requires further study. ...
Preprint
Full-text available
The phenomenon of high ice water content (HIWC) occurs in mesoscale convective systems (MCS) when a large number of small ice particles with typical sizes of a few hundred micrometers, concentrations of the order of 102–103 L-1 and IWC exceeding 1 g m-3 are found at high altitudes. HIWC regions in MCSs may extend vertically up to 10 km above the melting layer and horizontally up to hundreds of kilometers, filling large volumes of the convective systems. HIWC has great geophysical significance due to its effect on precipitation formation, the hydrological cycle, and the radiative properties of MCSs. It is also recognized as a hazard for commercial aviation operations since it can result in engine power loss and in the malfunctioning of aircraft data probes. This study summarizes observational and numerical simulation efforts leading to the development of a conceptual model for the production of HIWC in tropical MCSs based on the data collected during the HAIC-HIWC campaign in French Guiana in 2015. It is hypothesized that secondary ice production (SIP) in the vicinity of the melting layer plays a key role in the formation and sustainability of HIWC. In-situ observations suggest that the major SIP mechanism in the vicinity of the melting layer is related to the fragmentation of freezing drops (FFD). Both in-situ data and numerical simulations suggest that the recirculation of drops through the melting layer could lead to the amplification of SIP. However, laboratory measurements remain insufficient to support the accurate model representation of FFD conclusively. The proposed conceptual model and simulation results motivate further efforts to extend reproducible laboratory measurements.
... The critical value of the dimensionless energy (DE C ) is set as 0.2. According to James et al. (2021), the probability of any ice-containing droplet fragment in the splash (Φ(T )) is set as 0.3. The frozen fraction is f(T ). ...
Article
Full-text available
The charge structure in thunderstorms may be strongly affected by different secondary ice production (SIP) processes, but has not been well understood. In this study, the impacts of three SIP mechanisms on microphysics and electrification in a squall line are investigated using model simulation, including the rime‐splintering, ice‐ice collisional breakup, and shattering of freezing drops. The parameterization of the three SIP mechanisms, a noninductive and an inductive charging parameterization are implemented in the spectral bin microphysics. The results show that with SIP processes included, the modeled radar reflectivity is more consistent with observation. It is found that both the mass and concentrations of graupel/hail are enhanced by SIP processes, while the diameter decreases. The mixing ratio of ice/snow decreases due to the rime‐splintering, and increases in mixing ratio are due to the shattering of freezing drops. Particle charging is significantly affected by SIP, leading to a dipole structure of the total charge density, which includes a lower negative and an upper positive charge region. With both the noninductive and inductive charging considered, the charge carried by graupel/hail changes from negative to a bipolar structure, and the charge sign carried by ice/snow is inverted due to the SIP. The modeled lightning activity is enhanced by implementing all three SIP processes, while if only considering the rime‐splintering process, the flash rate would be suppressed. The insights obtained from this study highlight the importance of considering different mechanisms of SIP in modeling the charge structure and lightning activity in thunderstorms.
... According to Korolev and Leisner (2020), SIP might proceed according to the following mechanisms: (a) droplet fragmentation during freezing (Takahashi and Yamashita, 1977; 30 Wildeman et al., 2017;Lauber et al., 2018;Keinert et al., 2020;Kleinheins et al., 2021), (b) rime-splintering, (c) fragmentation during ice-ice particle collisions (Vardiman, 1978;Takahashi et al., 1995;Grzegorczyk et al., 2023), ice fragmentation due to (d) thermal shock (Dye and Hobbs, 1968;King and Fletcher, 1976b), (e) sublimation (Oraltay and Hallett, 1989;Dong et al., 1994;Bacon et al., 1998) and (f) activation of INPs in transient supersaturation (e.g., Prabhakaran et al., 2020). Yet another SIP mechanism occurring during the break-up of freezing droplets on impact with smaller ice particles, suggested by Phillips et al. 35 (2018), was supported experimentally by James et al. (2021). None of these proposed SIP mechanisms has been sufficiently characterized so far. ...
Preprint
Full-text available
Mixed-phase clouds are essential for Earth’s weather and climate system. Ice multiplication via secondary ice production (SIP) is thought to be responsible for the observed strong increase of ice particle number concentration in mixed-phase clouds. In this study, we focus on the rime-splintering also known as the Hallett-Mossop (HM) process, which still lacks physical and quantitative understanding. We report on an experimental study of rime-splintering conducted in a newly developed setup under conditions representing convective mixed-phase clouds in the temperature range of −4 °C to −10 °C. The riming process was observed with high-speed video microscopy and infrared thermography, while potential secondary ice particles (SI) in the super-micron size range were detected by a custom-build ice counter. Contrary to earlier HM experiments, where up to several hundreds of SI particles per mg rime were found at −5 °C, we found no evidence of productive SIP, which fundamentally questions the importance of rime-splintering. Further, we could exclude two potential mechanisms suggested as explanation for rime-splintering: freezing of droplets upon glancing contact with the rimer and fragmentation of spherically freezing droplets on the rimer surface. The break-off of sublimating fragile rime spires was observed to produce very few SI particles, insufficient to explain the large numbers of ice particles reported in earlier studies. In the transition regime between wet and dry growth, in analogy to phenomena of deformation of drizzle droplets upon freezing, we also observed formation of spikes on the rimer surface, which might be a source of SIP.
... F and G are the specific heat capacity of water and the specific latent heat of freezing, respectively. A'(& = 0.2, and Φ is 0.3 according to James et al. (2021). All ice fragments are assumed to be tiny in this mode. ...
Preprint
Full-text available
Ice microphysics controls cloud electrification in thunderstorms, and the various secondary ice production (SIP) processes are vital in generating high ice concentration. However, the role of SIP in cold-season thunderstorms is not well understood. In this study, the impacts of SIP on the electrification in a thunderstorm occurred in late November is investigated using model simulations. The parameterizations of three SIP processes are implemented in the model, including the rime-splintering, ice-ice collisional breakup, and shattering of freezing drops. In addition, a noninductive and an inductive charging parametrization, as well as a bulk discharging model are coupled with the spectral bin microphysics scheme. The results show the simulated storm intensity and temporal variation of flash rate are improved after SIP parametrizations are implemented in the model. Among the three SIP processes, the rime-splintering and shattering of freezing drops have stronger impacts on the storm than the ice-ice collisional breakup. The graupel and snow concentration are enhanced while their sizes are suppressed due to the SIP. The changes in the ice microphysics result in substantial changes in the charge structure. The total charge density changes from an inverted tripole structure to a dipole structure (tripole structure at some locations) after SIP is considered in the model, mainly due to the enhanced collision between graupel and ice, and riming at temperatures warmer than -20 °C. These changes lead to an enhancement of vertical electric field, especially in the mature stage, which explains the improved modelling of flash rate. The results highlight that the cold-season cloud electrification is very sensitive to the SIP.
Article
Multiple mechanisms have been proposed to explain secondary ice production (SIP), and SIP has been recognized to play a vital role in forming cloud ice crystals. However, most weather and climate models do not consider SIP in their cloud microphysical schemes. In this study, in addition to the default rime splintering (RS) process, two SIP processes, namely, shattering/fragmentation during freezing of supercooled rain/drizzle drops (DS) and breakup upon ice–ice collisions (BR), were implemented into a two-moment cloud microphysics scheme. Besides, two different parameterization schemes for BR were introduced. A series of sensitivity experiments were performed to investigate how SIP impacts cloud microphysics and cloud phase distributions in warm-based deep convective clouds developed in the central part of Europe. Simulation results revealed that cloud microphysical properties were significantly influenced by the SIP processes. Ice crystal number concentrations (ICNCs) increased up to more than 20 times and surface precipitation was reduced by up to 20% with the consideration of SIP processes. Interestingly, BR was found to dominate SIP, and the BR process rate was larger than the RS and DS process rates by four and three orders of magnitude, respectively. Liquid pixel number fractions inside clouds and at the cloud top decreased when implementing all three SIP processes, but the decrease depended on the BR scheme. Peak values of ice enhancement factors (IEFs) in the simulated deep convective clouds were 10 ² –10 ⁴ and located at −24°C with the consideration of all three SIP processes, while the temperature dependency of IEF was sensitive to the BR scheme. However, if only RS or RS and DS processes were included, the IEFs were comparable, with peak values of about 6, located at −7°C. Moreover, switching off the cascade effect led to a remarkable reduction in ICNCs and ice crystal mass mixing ratios. Significance Statement The cloud phase is found to have a significant impact on cloud evolution, radiative properties, and precipitation formation. However, the simulation of the cloud phase is a big challenge for cloud research because multiple processes are not well described or missing in numerical models. In this study, we implemented two secondary ice production (SIP) processes, namely, shattering/fragmentation during the freezing of supercooled rain/drizzle drops and breakup upon ice–ice collisions, which are missing in most numerical models. Sensitivity experiments were conducted to investigate how SIP impacts cloud microphysics and cloud phase in deep convective clouds. We found that SIP significantly impacts in-cloud and cloud-top phase distribution. We also identified that the collisional breakup of ice particles is the dominant SIP process in the simulated deep convective clouds.
Preprint
Full-text available
Representing single or multi-layered mixed-phase clouds (MPCs) accurately in global climate models (GCMs) is critical for capturing climate sensitivity and Arctic amplification. Ice multiplication, or secondary ice production (SIP), can increase the ice crystal number concentration (ICNC) in MPCs by several orders of magnitude, affecting cloud properties and processes. Here, we propose a machine-learning approach, called Random Forest SIP (RaFSIP), to parameterize the effect of SIP on stratiform MPCs. The RaFSIP scheme uses few input variables available in models and considers rime splintering, ice-ice collisional break-up, and droplet-shattering, operating at temperatures between 0 and -25 ˚C. The training dataset for RaFSIP was derived from two-year pan-Arctic simulations with the Weather Research and Forecasting (WRF) model with explicit representations of SIP processes. The RaFSIP scheme was evaluated offline against WRF simulation outputs, then integrated within WRF. The parameterization exhibits stable performance over a simulation year, and reproduced predictions of ICNC with explicit microphysics to within a factor of 3. The coupled WRF-RaFSIP scheme can replicate regions of enhanced SIP and accurately map ICNCs and liquid water content, particularly at temperatures above -10 ˚C. Uncertainties related to the RaFSIP representation of MPCs marginally affected surface cloud radiative forcing in the Arctic, with radiative biases of lower than 3 Wm-2 compared to simulations with explicit SIP microphysics. Training from a few high-resolution model grid points did not limit the predictive skill of RaFSIP, with the approach opening up new avenues for model simplification and process description in GCMs by physics-guided machine learning algorithms.
Preprint
Full-text available
We experimentally investigated collisions of supercooled water drops (∼ 5 mm in diameter) with ice particles of a similar size placed on a glass slide at temperatures T ≥ −12 °C. Our results showed that secondary drops were generated during both the spreading and retraction phase of the supercooled water drop impact. The secondary drops generated during the spreading phase were emitted too fast to quantify. However, quantification of the secondary drops generated during the retraction phase with diameters > 0.1 mm showed that 5–10 secondary drops formed per collision, with approximately 30 % of the secondary drops freezing over a temperature range of −4 °C ≤ T ≤ −12 °C. Our investigation provides the first dedicated laboratory study of collisions of supercooled water drops with ice particles as a secondary ice production mechanism. Our results suggest that this secondary ice production mechanism may be significant for ice formation in atmospheric clouds containing large supercooled drops and ice particles.
Article
Full-text available
Secondary ice production (SIP) plays a key role in the formation of ice particles in tropospheric clouds. Future improvement of the accuracy of weather prediction and climate models relies on a proper description of SIP in numerical simulations. For now, laboratory studies remain a primary tool for developing physically based parameterizations for cloud modeling. Over the past 7 decades, six different SIP-identifying mechanisms have emerged: (1) shattering during droplet freezing, (2) the rime-splintering (Hallett-Mossop) process, (3) fragmentation due to ice-ice collision, (4) ice particle fragmentation due to thermal shock, (5) fragmentation of sublimating ice, and (6) activation of ice-nucleating particles in transient supersaturation around freezing drops. This work presents a critical review of the laboratory studies related to secondary ice production. While some of the six mechanisms have received little research attention, for others contradictory results have been obtained by different research groups. Unfortunately, despite vast investigative efforts , the lack of consistency and the gaps in the accumulated knowledge hinder the development of quantitative descriptions of any of the six SIP mechanisms. The present work aims to identify gaps in our knowledge of SIP as well as to stimulate further laboratory studies focused on obtaining a quantitative description of efficiencies for each SIP mechanism .
Article
Full-text available
By 31 December 2018, 39 537 quality-controlled reports of large hail had been submitted to the European Severe Weather Database (ESWD) by volunteers and ESSL. This dataset and the NatCatSERVICE Database of Munich RE jointly allowed us to study the hail hazard and its impacts across Europe over a period spanning multiple decades. We present a spatiotemporal climatology of the ESWD reports, diurnal and annual cycles of large hail, and indicate where and how they may be affected by reporting biases across Europe. We also discuss which hailstorms caused the most injuries and present the only case with hail fatalities in recent times. Additionally, we address our findings on the relation between hail size to the type of impacts that were reported. For instance, the probability of reported hail damage to roofs, windows, and vehicles strongly increases as hail size exceeds 5 cm, while damage to crops, trees, and greenhouses is typically reported with hailstone diameters of 2–3 cm. Injuries to humans are usually reported with hail 4 cm in diameter and larger, and number of injuries increases with increasing hail size. Using the NatCatSERVICE data, we studied economic losses associated with hailstorms occurring in central Europe and looked for long-term changes. The trend in hail losses and the annual number of hail loss days since 1990 to 2018 are compared to that of meteorological conditions favorable for large hail as identified by ESSL’s Additive Regression Convective Hazards model. Both hail loss days and favorable environments show an upward trend, in particular since 2000.
Article
Full-text available
Over the decades, the cloud physics community has debated the nature and role of aerosol particles in ice initiation. The present study shows that the measured concentration of ice crystals in tropical mesoscale convective systems exceeds the concentration of ice nucleating particles (INPs) by several orders of magnitude. The concentration of INPs was assessed from the measured aerosol particle concentration in the size range of 0.5 to 1 µm. The observations from this study suggest that primary ice crystals formed on INPs make only a minor contribution to the total concentration of ice crystals in tropical mesoscale convective systems. This is found by comparing the predicted INP number concentrations with in situ ice particle number concentrations. The obtained measurements suggest that ice multiplication is the likely explanation for the observed high concentrations of ice crystals in this type of convective system.
Article
Full-text available
Aircraft measurements of two cumulus clouds were made during the Ice and Precipitation Initiation in Cumulus campaign over the British Isles. The 18 May 2006 cloud had high concentrations of ice particles and conditions were conducive for the Hallett-Mossop (HM) process of secondary ice production, but the 13 July 2005 cloud had low concentrations. A bin-resolved cloud model was used to investigate several factors that are known to control the HM process using the observations of the two clouds. For the 2006 cloud, the model results show that the fast production of graupel by directly freezing of supercooled raindrops through collisional collection with ice particles was crucial to the activation of the HM process. Switching-off raindrop freezing led to much delayed and suppressed formation of graupel particles, and hence, a negligible HM process. Sensitivity studies were performed on the concentration of primary ice particles required to kick-start the HM process. It was found that a concentration of the first ice as low as 0.01 L− 1 could be sufficient, as long as there was a large enough concentration of cloud droplets (small and large) available when a significant number of graupel particles developed in the HM temperature zone. For the modelled 2005 cloud, the HM process did not operate effectively mainly because of the low concentration of supercooled raindrops and hence graupel. The HM process was also hindered by the relatively greater number of aerosols, and higher temperatures at cloud base and top.
Article
Full-text available
Measured ice crystal concentrations in natural clouds at modest supercooling (~>-10°C) are often orders of magnitude greater than the number concentration of primary ice nucleating particles. Therefore, it has long been proposed that a Secondary Ice Production process must exist that is able to rapidly enhance the number concentration of the ice population following initial primary ice nucleation events. Secondary Ice Production is important for the prediction of ice crystal concentration and the subsequent evolution of some types of clouds, but the physical basis of the process is not understood and the production rates are not well constrained. In November 2015 an international workshop was held to discuss the current state of the science and future work to constrain and improve our understanding of secondary ice production processes. Examples and recommendations for in-situ observations, remote sensing, laboratory investigations and modelling approaches are presented.
Article
Drop splash and breakup on cylindrical surfaces play an important role in a wide variety of industrial applications. In this work, water drops with a wide range of impact velocities (1.4 m/s–4.5 m/s) and cylindrical stainless steels with different diameters (1 mm–20 mm) are employed to investigate the asymmetric splash and breakup characteristics of drops impacting on cylindrical superhydrophobic surfaces. We identify two interesting phenomena, asymmetric splash and converging breakup. The splash behavior is found to be asymmetric in different directions, and the drops preferentially splash in the axial direction. Fundamentally, we propose two disparate splash thresholds, referring to the Weber number We and the diameter ratio D* = D/D0, in the azimuthal and axial directions, respectively. The converging breakup is caused by the much more rapid converging of the liquid rim in the axial direction than in the azimuthal direction. The aspect ratio βzmax/βxmax, governing the converging breakup, increases with We and decreases with D*. Fortuitously, the splashing angle θ is demonstrated to only depend on D* rather than We, and the relational expression of θ and D* is provided. Ultimately, we put forward universal relations between the mean diameter and velocity of secondary droplets, resulting from the converging breakup, and the dimensionless parameter We/D*. The results of this work are expected to provide valuable insights into anti-icing and microfluidics fields.
Article
A numerical formulation is provided for secondary ice production during fragmentation of freezing raindrops or drizzle. This is obtained by pooling laboratory observations from published studies and considering the physics of collisions. There are two modes of the scheme: fragmentation during spherical drop freezing (mode 1) and during collisions of supercooled raindrops with more massive ice (mode 2). The empirical scheme is for atmospheric models. Microphysical simulations with a parcel model of fast ascent (8 m s⁻¹) between -10° and -20°C are validated against aircraft observations of tropical maritime deep convection. Ice enhancement by an order of magnitude is predicted from inclusion of raindrop-freezing fragmentation, as observed. The Hallett-Mossop (HM) process was active too. Both secondary ice mechanisms (HM and raindrop freezing) are accelerated by a positive feedback involving collisional raindrop freezing. An energy-based theory is proposed explaining the laboratory observations of mode 1, both of approximate proportionality between drop size and fragment numbers and of their thermal peak. To illustrate the behavior of the scheme in both modes, the glaciation of idealized monodisperse populations of drops is elucidated with an analytical zero-dimensional (0D) theory treating the freezing in drop-ice collisions by a positive feedback of fragmentation. When drops are too few or too small (≪1 mm), especially at temperatures far from -15°C (mode 1), there is little raindrop-freezing fragmentation on realistic time scales of natural clouds, but otherwise, high ice enhancement (IE) ratios of up to 100-1000 are possible. Theoretical formulas for the glaciation time of such drop populations, and their maximum and initial growth rates of IE ratio, are proposed.
Article
The present study is devoted to the experimental investigation and theoretical modelling of the interaction between fluid flow and solidification during the impact of supercooled water drops onto an ice surface. Using a high-speed video system, the impact process is captured with a high spatial and temporal resolution in a side view. The lamella thinning and the residual ice layer thickness in the centre of impact are determined from the high-speed videos for varying drop and surface temperatures, and impact velocities. It is shown that the temperature of the impact surface has a negligible influence and the drop temperature has a dominating influence on the lamella thinning and the final ice layer thickness. For decreasing drop temperatures, higher freezing rates cause a decreased rate of lamella thinning and a larger thickness of the resulting ice layer. On the other hand, a higher impact velocity causes an increasing speed of lamella thinning and a smaller thickness of the resulting ice layer. Based on a postulated flow in the spreading lamella and considering the ice layer growth and the developing viscous boundary layer, the upper limit for the resulting ice layer thickness is theoretically modelled. The theory shows very good agreement with the experimental results for all impact conditions. Based on the derived theoretical scaling, a semi-empirical equation is obtained which allows an a priori prediction of the final ice layer thickness resulting from a single drop impact, knowing the impact conditions. This capability is important for the improvement of existing ice accretion models.
Article
Ice particle formation in tropospheric clouds significantly changes cloud radiative and microphysical properties. Ice nucleation in the troposphere via homogeneous freezing occurs at temperatures lower than -38 °C and relative humidity with respect to ice above 140%. In the absence of these conditions, ice formation can proceed via heterogeneous nucleation aided by aerosol particles known as ice nucleating particles (INPs). In this paper we describe new developments in identifying the heterogeneous freezing mechanisms, atmospheric relevance, uncertainties and unknowns about INPs. We explain how conventional wisdom regarding the requirements of INPs is changing as new studies discover physical and chemical properties of these particles. INP sources and known reasons for their ice nucleating properties are presented. We also highlight where more studies are required to systematically identify particle properties that facilitate ice nucleation. The atmospheric relevance of long range transport, aerosol aging and coating studies (in the laboratory) of INPs are also presented. Possible mechanisms for processes that change the ice nucleating potential of INPs and the corresponding challenges in understanding and applying these in models are discussed. How primary ice nucleation affects total ice crystal number concentrations in clouds and the discrepancy between INP concentrations and ice crystal number concentrations are presented. Finally, we discuss limitations of parameterizing INPs and of models in representing known and unknown processes related to heterogeneous ice nucleation processes.