Eco Ma 1

Embed Size (px)

Citation preview

  • 8/3/2019 Eco Ma 1

    1/9

    Measurements of meteor smoke particles during the ECOMA-2006campaign: 1. Particle detection by active photoionization

    Markus Rapp, Irina Strelnikova

    Department of Radar Soundings and Sounding Rockets, Leibniz-Institute of Atmospheric Physics, University of Rostock, Schlosstr. 6, 18225 Kuhlungsborn, Germany

    a r t i c l e i n f o

    Article history:

    Accepted 2 June 2008Available online 12 June 2008

    Keywords:

    Meteor smoke particles

    In situ measurements

    a b s t r a c t

    We present a new design of an in situ detector for the study of meteor smoke particles (MSPs) in the

    middle atmosphere. This detector combines a classical Faraday cup with a xenon-flashlamp for the

    active photoionization/photodetachment of MSPs and the subsequent detection of corresponding

    photoelectrons. This instrument was successfully launched in September 2006 from the Andya Rocket

    Range in Northern Norway. A comparison of photocurrents measured during this rocket flight and

    measurements performed in the laboratory proves that observed signatures are truly due to

    photoelectrons. In addition, the observed altitude cut-off at 60 km (i.e., no signals were observed

    below this altitude) is fully understood in terms of the mean free path of the photoelectrons in the

    ambient atmosphere. This interpretation is also proven by a corresponding laboratory experiment.

    Consideration of all conceivable species which can be ionized by the photons of the xenon-flashlamp

    demonstrates that only MSPs can quantitatively explain the measured currents below an altitude of

    90 km. Above this altitude, measured photocurrents are most likely due to photoionization of nitric

    oxide. In conclusion, our results demonstrate that the active photoionization and subsequent detection

    of photoelectrons provides a promising new tool for the study of MSPs in the middle atmosphere.

    Importantly, this new technique does not rely on the a priori charge of the particles, neither is the

    accessible particle size range severely limited by aerodynamical effects. Based on the analysis described

    in this study, the geophysical interpretation of our measurements is presented in the companion paperby Strelnikova, I., et al. [2008. Measurements of meteor smoke particles during the ECOMA-2006

    campaign: 2. results. Journal of Atmospheric and Solar-Terrestrial Physics, this issue, doi:10.1016/

    j.jastp.2008.07.011].

    & 2008 Elsevier Ltd. All rights reserved.

    1. Introduction

    Meteor smoke particles (MSPs) are thought to be formed by

    the recondensation of metal- and silicon-containing molecules

    originating from the ablation of meteoroids in the 70120 km

    altitude range (Plane, 2003). Starting with the 1960s, modelstudies suggested that MSPs should exist in the mesosphere/lower

    thermosphere with number densities of up to 104 particles=cm3

    and corresponding radii ranging from the sub-nanometer range to

    a few nanometers (Rosinski and Snow, 1961; Hunten et al., 1980).

    Despite these tiny dimensions, it has been suggested that MSPs

    are involved in a variety of atmospheric processes such as the

    nucleation of mesospheric ice clouds, heterogeneous chemistry,

    and the formation of nitric acid trihydrate (NAT)-particles in polar

    stratospheric clouds which are involved in ozone destruction in

    the polar spring (e.g., Rapp and Thomas, 2006; Summers and

    Siskind, 1999; Voigt et al., 2005). Despite obvious interest in MSPs,

    measurements of such particles have proven difficult so that today

    only a few measured altitude profiles are available from charged

    particle measurements on sounding rockets (Schulte and Arnold,

    1992; Gelinas et al., 1998; Horanyi et al., 2000; Croskey et al.,2001; Lynch et al., 2005; Rapp et al., 2005; Amyx et al., 2008 ) and

    from incoherent scatter radar experiments (Rapp et al., 2007;

    Strelnikova et al., 2007).

    Most of these earlier in situ measurements used Faraday cup-

    based instruments based on the original design of Havnes et al.

    (1996) who was the first to detect charged (ice) particles in the

    polar summer mesopause region. For this type of detector,

    however, it has become clear in the meantime that it possesses

    a severely limited detection efficiency for the smallest MSPs

    (i.e., below$12 nm) and for which the smallest detectable radius

    even varies with altitude as a consequence of aerodynamics

    (Horanyi et al., 1999; Rapp et al., 2005; Hedin et al., 2007).

    In addition, it has recently been speculated that such

    ARTICLE IN PRESS

    Contents lists available at ScienceDirect

    journal homepage: www.elsevier.com/locate/jastp

    Journal ofAtmospheric and Solar-Terrestrial Physics

    1364-6826/$- see front matter& 2008 Elsevier Ltd. All rights reserved.doi:10.1016/j.jastp.2008.06.002

    Corresponding author. Tel.: +49 3829368200; fax: +49 382936850.

    E-mail address: [email protected] (M. Rapp).

    Journal of Atmospheric and Solar-Terrestrial Physics 71 (2009) 477485

    http://dx.doi.org/10.1016/j.jastp.2008.07.011http://dx.doi.org/10.1016/j.jastp.2008.07.011http://www.sciencedirect.com/science/journal/atphttp://www.elsevier.com/locate/jastphttp://dx.doi.org/10.1016/j.jastp.2008.06.002mailto:[email protected]:[email protected]://dx.doi.org/10.1016/j.jastp.2008.06.002http://www.elsevier.com/locate/jastphttp://www.sciencedirect.com/science/journal/atphttp://dx.doi.org/10.1016/j.jastp.2008.07.011http://dx.doi.org/10.1016/j.jastp.2008.07.011
  • 8/3/2019 Eco Ma 1

    2/9

    measurements which rely on the impact of particles on the

    detector electrode may be influenced by secondary charging

    effects such as fragmentation and triboelectric charging (Barjatya

    and Swenson, 2006; Havnes and Naesheim, 2007; Amyx et al.,

    2008). Measurements with incoherent scatter radars, on the other

    hand, have so far only proven to be feasible with the most

    powerful incoherent scatter radar, i.e., the Arecibo radar, whereas

    measurements with even the EISCAT UHF radar were onlyindicative of the existence of MSPs (Rapp et al., 2007). In addition,

    even with the Arecibo radar, measurements could only be

    performed at altitudes with sufficient D-region ionization, i.e., in

    daytime and above $85km (Strelnikova et al., 2007). Hence, it is

    obvious that new measurements which should cover a larger

    altitude range and which should be able to detect MSPs of all sizes

    are highly desirable.

    In the current manuscript, we present the concept and first

    results of a new rocket borne particle detector which uses active

    photoionization to measure the concentration of MSPs and hence

    largely avoids limitations from aerodynamics and secondary

    charging effects and does not depend on the a priori charge of

    the particles. In Section 2 we introduce the basic design of this

    detector. Initial results from a rocket flight during the ECOMA-2006 campaign (ECOMA existence and charge state of meteor

    smoke particles in the middle atmosphere) are presented in

    Section 3. These flight data are then discussed in the scope

    of laboratory measurements in Section 4.1 and with respect to

    contamination from other ionizable species (like nitric oxide) in

    Section 4.2. The geophysical results from the discussed rocket

    flight are presented in the companion paper by Strelnikova et al.

    (2008).

    2. Instrument description

    The particle detector applied for the current study is a

    combination of a Faraday cup and a xenon-flashlamp for the

    photoionization of particles. A photo of the instrument together

    with a schematic of the detector is shown in Fig. 1.

    The design of the Faraday cup itself is very close to the original

    design ofHavnes et al. (1996) who were the first to detect charged

    nanoparticles (in their case consisting of water ice) in the

    mesosphere. The Faraday cup comprises a collector electrode

    (held at payload potential by the negative feedback loop of the

    electrometer) for the measurement of particles of either positive

    or negative charge and two shielding grids (biased at 6:2 V

    relative to payload potential) to shield the collector electrode from

    ambient thermal electrons and ions. Note that unlike these

    thermal electrons and ions the particles have a much larger mass

    such that their kinetic energy (i.e., 12 mpv2r where mp is the particle

    mass and vr is the rocket velocity) is large enough to penetrate thepotential barrier set by the shielding grids.

    The novelty of this instrument, however, is the combination of

    the Faraday cup design with a xenon-flashlamp. The flashlamp is a

    commercially available lamp (Perkin Elmer FX1162, see http://

    optoelectronics.perkinelmer.com for technical details) which we

    operate at a repetition rate of 20 Hz with an energy per flash of

    0.5 J. The broadband spectrum of each flash contains a sufficiently

    large number of UV photons down to a minimum wavelength of

    $110 nm corresponding to a maximum photon energy of 11.3 eV

    (see Fig. 2). The UV photons may now create photoelectrons by

    photoionization and/or photodetachment of an electron from a

    neutral/negatively charged species with sufficiently low work

    function (or electron affinity) which may in turn be detected at

    the detector electrode as a short charge pulse. We note thatinstruments employing lamps to actively ionize atmospheric

    constituents were applied earlier (e.g., Croskey et al., 2003, and

    references therein).

    In order to be able to record both the continuous charge signal

    from the naturally charged MSPs which are large enough to

    penetrate to the detector electrode and the very short charge

    pulses excited by the UV flash, we have chosen to operate the

    electrometer in the so-called integrate and dump mode. This has

    the advantage that even the smallest charge signatures are not

    missed provided the noise level of the electrometer is sufficiently

    low. This means that the electrometer is in fact operated as an

    integrator which has to be discharged shortly before a flash. The

    charge pulse created by the flash is then integrated by an inverting

    amplifier (i.e., a positive charge leads to a negative slope registered

    by the integrator) and the resulting signal is digitized by a 16 bit

    analog-to-digital converter. In addition to the charge pulse, the

    amplifier also samples any existent DC-current due to naturally

    charged particles of either polarity. In order to properly sample the

    shape of the UV-created charge pulses on the one hand and tomeasure the rather small DC-currents due to naturally charged

    particles on the other hand, we sample the output signal of the

    integrator at two sampling frequencies: the charge pulsewhich

    originates from an effective ionized volume between 2.5cm

    (photoelectrons from closer distances are actually in the shadow

    of the flashlight itself and cannot reach the electrode) and 70 cm

    upstream of the detectorhas a typical duration ofo0:5ms and is

    sampled with 100 kS/s (=kilosamples/second), stored in a memory,

    and subsequently transmitted to ground by telemetry in non-real

    time (note, however, that the start time of each such data dump is

    exactly defined). DC-currents, however, are sampled with 1 kS/s

    and transmitted to ground in real time (note that the first sample

    of this signal also contains the charge pulse, however, at a much

    poorer temporal resolution). If the integrator reaches its positive ornegative maximum value it is automatically discharged such that

    ARTICLE IN PRESS

    Fig. 1. Photo and schematic of the ECOMA-particle detector. See text for details.

    M. Rapp, I. Strelnikova / Journal of Atmospheric and Solar-Terrestrial Physics 71 (2009) 477485478

    http://optoelectronics.perkinelmer.com/http://optoelectronics.perkinelmer.com/http://optoelectronics.perkinelmer.com/http://optoelectronics.perkinelmer.com/
  • 8/3/2019 Eco Ma 1

    3/9

    this amplifier setup provides the capability to continuously

    measure a broad current range, i.e., from the sub-picoamp to the

    0:5mA range. Finally, after 50 ms (i.e., shortly before the next flash)the next discharge of the integrator is initiated and the whole

    sequence starts from the beginning.

    As applied to the measurement of MSPs, the conceptual idea

    behind our particle detector may be summarized as follows: In

    Fig. 3 we have sketched the particle size distribution of MSPs as

    predicted by microphysical models (Hunten et al., 1980; Megner

    et al., 2006). In this size distribution, particles with radii larger

    than a minimum radius (here indicated by the vertical blue line)

    may be directly detected by particle impact and subsequent

    charge deposition at the collector electrode of the Faraday cup.

    Particles with radii smaller than this minimum radius may not be

    detected because of aerodynamic effects on their trajectories

    (e.g., Rapp et al., 2005; Hedin et al., 2007). All of the particles,

    however, may be photoionized and corresponding photoelectrons

    may be detected as a charge pulse. Hence, the particle detector

    provides information on the number density of charged particles

    which are larger than a certain threshold size and on the total

    number density of all MSPs.

    Note that first results of conventional Faraday cup measure-

    ments with this instrument from a rocket flight in October 2004were already presented in Rapp et al. (2005). At that time,

    however, the photoelectron measurements with the detector

    were impaired by electromagnetic interference from the flash-

    electronics and could not be used. In the present manuscript, we

    will focus on new measurements obtained in September 2006

    during which photoelectron measurements yielded high quality

    data. In the following section we will present actual data from this

    rocket flight with which we will demonstrate the feasibility of the

    detector concept described above.

    3. Atmospheric measurements

    In September 2006, the ECOMA-2006 campaign was conductedfrom the North-Norwegian Andya Rocket Range 69N. ECOMA

    ( existence and charge state of MSPs in the middle atmosphere)

    is an international research program led by the Leibniz Institute of

    Atmospheric Physics in Germany and the Norwegian Defence

    Research Establishment in Norway (with additional contributions

    from Austria, Sweden, and the USA) dedicated to the study of

    MSPs and their atmospheric environment. Besides the particle

    detector described in the preceding section, the payload carried

    instruments to characterize the D-region plasma, a particle

    sampler for the in-flight collection of MSPs, and instruments

    to measure neutral air density, temperature, and turbulence.

    The atmospheric background state was monitored by means of

    the EISCAT UHF and VHF radars, the ALOMAR RMR lidar, and the

    ALOMAR sodium lidar. An overview of all measurements obtained

    during the campaign along with a more detailed description of all

    involved instruments is presented in the companion paper by

    Strelnikova et al. (2008). In this manuscript, however, our focus is

    on the flight performance of the ECOMA particle detector.

    The ECOMA-01 payload was launched on September 8, 2006 at

    22:17:00 UT and reached an apogee of 130.6km. During the entire

    rocket flight the particle detector worked nominally and provided

    data in both data channels. In order to demonstrate the quality of

    these measurements we present an arbitrarily chosen raw data

    sample from about 83:95 0:05 km on the upleg part of the rocketflight. In Fig. 4, the black symbols and lines show the charge signal

    from the integrator circuit in digital units (du). The red symbols

    and lines are corresponding currents derived by taking the time

    derivative of the integrator output and taking into account that

    1 du corresponds to a charge of 0.18 fC. Blue vertical lines indicate

    the times when the integrator was discharged after which a flash

    was triggered. The upper panel shows results from the 1 kHz data

    channel which primarily records charge signatures and currents

    due to the impact of particles on the collector electrode. In the

    lower panel, corresponding signals from the fast (100 kHz) data

    channel are shown which clearly show the pulse due to

    photoelectrons with an amplitude of $ 4:5n A on top of a

    smooth DC-background, in this case of about 50 pA. Note that

    the photoelectron current due to the flash is certainly alsorecorded in the slow data channel, however, due to its poorer

    ARTICLE IN PRESS

    Fig. 2. Spectrum of the xenon-flashlamp in the UV wavelength range measured with a plane grating vacuum spectrometer. Note that the cut-off at 110nm is due to the

    transmission properties of the window of the flashlamp.

    M. Rapp, I. Strelnikova / Journal of Atmospheric and Solar-Terrestrial Physics 71 (2009) 477485 479

  • 8/3/2019 Eco Ma 1

    4/9

    time resolution, its peak value is not captured (though the

    recorded charge signal is certainly the same).From the raw data of the type described above, we have now

    determined the background DC-currents from the slow data

    channel (where the first sample after each flash which contains

    the additional charge due to photoelectrons has been omitted)

    and the maximum current from each photoelectron pulse due to

    the xenon-flash from the fast data channel. Corresponding

    altitude profiles are presented in Fig. 5 for both the upleg and

    downleg part of the rocket flight. The background measurements

    of the particle charge (left panel in Fig. 5) reveal a prominent

    negative current layer in the altitude range between 80 and 90 km

    indicative of negatively charged particles. This layer is qualita-

    tively very similar to the previous measurements by Faraday cup

    instruments reported, e.g., by Lynch et al. (2005) and Rapp et al.

    (2005). A striking difference to the Rapp et al. (2005) results is,however, the opposite polarity of the particle charge. As discussed

    ARTICLE IN PRESS

    Fig. 3. Sketch of the particle size distribution of meteor smoke particles as predicted by microphysical models ( Hunten et al., 1980; Megner et al., 2006). The blue vertical

    line indicates that only a part of the size distribution may be detected by direct particle impact in a particle detector due to aerodynamical effects on the trajectories ofsmall particles (e.g., Rapp et al., 2005; Hedin et al., 2007). All of the particles, however, may be photoionized and corresponding photoelectrons may be detected as a charge

    pulse.

    Fig. 4. Raw data samples for the slow (i.e., 1 kHz, upper panel) and fast data

    channel (i.e., 100kHz, lower panel) from an altitude of $83:95km on the upleg

    part of the rocket trajectory. The blacksymbols and lines show the charge signal

    from the integrator circuit in digital units (du). The red symbols and lines are

    corresponding currents derived by taking the time derivative of the integrator

    output. Blue vertical lines indicate the times when the integrator is discharged

    after which the flash is triggered. Note that a positive slope in the integrator output

    indicates a negative current, in this case of about 50 pA. Note further that the

    photoelectron current due to the flash is also recorded in the slow data channel,however, due to its poorer time resolution, its peak value is not captured.

    Fig. 5. Overview of current measurements from the ECOMA flight on September 8,

    2006. Left panel: current measurements due to direct particle impacts on the

    electrode recorded as a DC-current in the slow data channel. Right panel:

    corresponding peak photoelectron currents recorded in the fast data channel.

    Black lines are for upleg measurements, red lines are for downleg.

    M. Rapp, I. Strelnikova / Journal of Atmospheric and Solar-Terrestrial Physics 71 (2009) 477485480

  • 8/3/2019 Eco Ma 1

    5/9

    in detail in Strelnikova et al. (2008), the major difference between

    these two flights is the much larger background ionization and a

    different electrode material in 2006 which might have caused

    potential differences in triboelectric charging.

    Above the negative layer, the current turns positive and even

    shows a very prominent positive peak at $93km which is the

    altitude where the EISCAT radars observed a sporadic E-layer

    (see Strelnikova et al., 2008, for details). It is interesting to notethat Horanyi et al. (2000) observed a similar morphology with a

    broad negative layer below a narrow positive layer which they

    interpreted as a sudden E-layer related to the deposition of

    meteoric metal ions. In our case, however, the positive peak is also

    seen on the downleg of the rocket flight when the instrument is

    not facing the ram (and no particles can enter the detector

    volume). Hence, we conclude that these currents must be due to

    contamination by positive ions. These measurements will be

    further discussedfor example with respect to the observed

    particle charge and triboelectric effectsin the companion

    paper by Strelnikova et al. (2008). Here, we now focus

    on the measurements due to the flash of the detector as shown

    in the right panel of Fig. 5. On upleg, large currents of up to

    10 nA were recorded above an altitude of$80 km. This currentdecays to about 2 nA at apogee but never returns to zero. On

    downleg, the measured current shows a similar behavior with

    slightly enhanced values from apogee down to 80 km. The largest

    difference, however, is that on downleg the (negative) current

    further increases with decreasing altitude until it reaches a

    maximum value of 15nA at $75km and disappears at about

    60 km, i.e., about 20 km lower than on upleg.

    In the next section we will critically discuss the following

    questions:

    (1) Is it possible to prove that the recorded currents in the fast

    data channel (called photo-currents hereafter) are indeed due

    to photoelectrons?

    (2) Is it possible to understand the differences of the recorded

    photo-currents on upleg and downleg?

    (3) And finally: Is it possible to show that the observed photo-

    currents originate from MSPs particles and not from other

    ionizable species such as nitric oxide, excited oxygen, or metal

    atoms?

    4. Discussion

    4.1. Laboratory experiments

    In order to answer questions 1 and 2 raised in Section 3 we

    devised a laboratory experiment as sketched in Fig. 6. The ECOMA

    particle detector was mounted inside a large vacuum chamber

    which was evacuated to a pressure below 104 mbar. Opposite thedetector we mounted a conical target coated with an aluminum

    surface shaped such that flash photons emitted by the xenon-

    flashlamp of the detector were not reflected back towards the

    collector electrode. In consequence, currents due to unwanted

    photoemission from the collector electrode could be avoided. UV

    photons from the xenon-flash, however, were sufficiently ener-

    getic (maximum energy 11.3 eV, see Section 2) to create photo-

    electrons at the surface of the aluminum target (workfunction

    4.2eV), part of which were then flying towards the collector

    electrode. A typical current pulse recorded during such measure-

    ments along with a current pulse recorded during the rocket flight

    in September 2006 is presented in Fig. 7. Obviously, both pulses,

    i.e., from the laboratory experiment and from the rocket flight,

    have exactly the same shape (note that the coincidence inamplitude is certainly by chance and is not meant to imply that

    the two pulses have exactly the same origin) which indicates that

    the pulses recorded during the rocket flight are indeed photo-

    electrons emitted by an ionizable species (see Section 4.2 below

    for arguments that these species must have been MSPs in most of

    the altitude range where photo-currents were observed). Note

    further that we were able to verify that the signals recorded in the

    laboratory were due to photoelectrons by varying a retarding

    potential between the target and the collector electrode.

    As expected for photoelectrons, the signal dropped to zero when

    the retarding potential reached a value of $ 6 V, i.e., at the

    expected kinetic energy of the photoelectrons ( flash energyminus work function). Note further that we also studied the effect

    of the potentials of the shielding grids on the collection of the

    photoelectrons. This showed that the potentials of the two grids

    only had a negligible effect on the collection of the photoelectrons

    indicating that the rejection of electrons in the eV energy range by

    this grid type is poor. Note, however, that another experiment in

    which we tested the rejection of thermal electrons (energy

    o0:1 eV) in a plasma chamber revealed that the rejection

    efficiency of these was close to 100%.

    Finally, we used the laboratory setup described above to study

    the pressure dependence of the recorded photocurrents. I.e., we

    varied the stationary pressure in the vacuum chamber between

    104 and 1 mbar and measured the maximum current due to

    photoelectrons emitted from the target. The results from thisexperiment are shown in the upper panel of Fig. 8. This indicates

    ARTICLE IN PRESS

    Fig. 6. Setup of laboratory measurements to identify the origin (and other

    properties) of the recorded currents due to the xenon-flash. The ECOMA-particle

    detector is brought into a vacuum chamber and flashes at a conical target coated

    with aluminum. The electrode voltage with respect to the vacuum chamber was

    varied between 8 V in order to determine the kinetic energy of the recorded

    photoelectrons (retarding field method).

    Fig. 7. Comparison of recorded current pulses due to photoelectrons produced bythe Xe-flash from the actual rocket flight (red curve) and laboratory measurements

    (black curve).

    M. Rapp, I. Strelnikova / Journal of Atmospheric and Solar-Terrestrial Physics 71 (2009) 477485 481

  • 8/3/2019 Eco Ma 1

    6/9

    that for pressures below $102 mbar, all photoelectrons can reach

    the collector electrode and are not affected by collisions with gas

    molecules. At higher pressures, however, the current due to

    photoelectrons drops until at $0:1 mbar no more photoelectrons

    can be detected. The reason for this behavior can be understood

    by considering the lower panel in Fig. 8. This figure shows

    calculations of the mean free path of an electron in air as a

    function of pressure. A pressure of 102 mbar corresponds to a

    mean free path of $5cm, which is about the dimension of our

    particle detector. For lower pressures, the mean free path of a

    photoelectron is hence sufficiently large such that it can be

    detected in our instrument without undergoing collisions with

    gas molecules. For higher pressures, however, collisions with gas

    molecules lead to the loss of a part of the photoelectrons until at a

    pressure of 0.1 mbar the mean free path has dropped by about a

    factor of 10 and no more photoelectrons can reach the collector

    electrode.

    We now come back to the results from our rocket flight,

    showing that on upleg the minimum altitude for photoelectrondetection was 80 km, whereas it decreased to 60 km on downleg.

    Taking into account the laboratory measurements shown in Fig. 8,

    this can be explained as follows: On upleg, the instrument is

    facing the ram such that the pressure inside the instrument is

    enhanced because of the aerodynamical compression of the

    ambient air. Monte Carlo simulations for a cup-like detector

    geometry as well as calculations using the RayleighPitot formula

    imply that for a rocket flight with an apogee of 130km the

    pressure should increase by about a factor of 13 at an altitude of

    80km (see Fig. 7 in Rapp et al., 2001). Using the densities and

    temperatures for the beginning of September from the climatol-

    ogy of Lubken (1999) this implies that at 80 km, the pressure

    should have been 13 102 mbar 0:13 mbar. Comparing this

    pressure to the pressure for which our laboratory measurementsshow that the photoelectron-current should have dropped to zero,

    we find that laboratory and flight results are in excellent

    agreement. Note that we may ignore the effect of the rather weak

    compression in the first few centimeters upstream of the detector

    on the creation of photoelectrons because the first 2.5 cm of the

    ionization volume are actually in the shadow of the flashlight such

    that photoelectrons from this volume can never reach the detector

    electrode (see Hedin et al., 2007 and our discussion of the

    effective ionization volume above). As for the downleg measure-ments, the Lubken (1999)-climatology predicts a pressure of

    0.24mbar at 60 km altitude. During this part of the rocket flight,

    however, the particle detector is in the wake, for which Monte

    Carlo simulations by Gumbel (2001) predict that the pressure

    should be reduced by about a factor of 2. Hence, on downleg,

    the current due to photoelectrons vanishes at a pressure of about

    0:5 0:24mbar 0:12 mbar, which is again almost exactly the

    pressure predicted by our laboratory experiments. Hence, the

    difference in the photoelectron measurements between upleg

    and downleg can be completely understood on the basis of the

    pressure dependence of the mean free path of the photoelectrons

    in the environment of the particle detector and its modification by

    the aerodynamics of the rocket flight. Interestingly, this analysis

    shows that the altitude range over which our method is capable ofdetecting signatures from ionizable species like meteor smoke

    particles is enhanced by$20 km on the downleg part of the rocket

    flight when the detector is in the wake.

    4.2. Ionization of other species than MSPs

    After we have demonstrated that the photocurrents recorded

    in our fast data channel are indeed due to photoelectrons excited

    by the photons of the xenon-flash, we now turn to the question if

    these photoelectrons truly originate from MSPs or from some

    other species.

    In order to answer this question, a survey of middle atmo-

    spheric species and corresponding threshold energies for either

    photoionization of a neutral molecule or an atom (i.e., theionization potential), or for the photodetachment of an electron

    from a negative ion (i.e., the electron affinity) reveals a list of

    candidates which we have summarized in Table 1. These

    candidates are nitric oxide (NO), excited molecular oxygen

    O21Dg, the metal atoms Fe and Na, and the negative ions NO

    3 ,

    CO3 , and O2 . Note that also other metal atoms like K, Ca, Mg, and

    Si have sufficiently low ionization potentials, however, for the

    calculations below we restrict ourselves to the case of Fe and Na

    which are by far the most abundant metal atoms in altitude range

    of interest (e.g., Plane, 2003).

    In order to judge whether the recorded photocurrents could

    be due to these species, we have tried to estimate the

    expected currents based on altitude profiles of their concentra-

    tions (see Fig. 9) and their properties with respect to photo-ionization/photodetachment (see Table 1). NO number densities

    have been taken from the HALOE-climatology of Siskind et al.

    (1998), O21Dg number densities from the rocket measurements

    by Gumbel et al. (1998), metal atom profiles are from Plane

    (2003), and number densities of negative ions are taken from the

    model study of Thomas and Bowman (1985). As for the meteor

    smoke particles, hypothetical profiles of number densities of MSPs

    of different sizes have been taken from the model study by

    Megner et al. (2006). References for corresponding photoioniza-

    tion cross sections and threshold energies are provided in Table 1.

    The photoionization/photodetachment cross sections of MSPs

    with radius rp at photon wavelength l are estimated using Mie-

    theory, i.e.,

    sMSPphotorp;l pr2p Qabsrp;l; nl; kl Y (1)

    ARTICLE IN PRESS

    Fig. 8. Upper panel: pressure dependence of the photoelectron current recorded in

    the laboratory (black symbols with blue error bars). The red lines are drawn to

    guide the eye. Lower panel: results of the calculation of the mean free path of a

    photoelectron in air as a function of air pressure.

    M. Rapp, I. Strelnikova / Journal of Atmospheric and Solar-Terrestrial Physics 71 (2009) 477485482

  • 8/3/2019 Eco Ma 1

    7/9

    where Qabs is the Mie absorption efficiency which we calculated

    using the publicly available Mie-code from the text book byBohren and Huffman (1983), n and k are the real and imaginary

    parts of the refractive index of the MSP material, and Y is the

    quantum yield for photoemission/photodetachment. Guided by

    the experimental finding that the yields of nanoparticles may

    reach very large values which are up to three orders of magnitude

    larger than the yields for the corresponding bulk material

    (e.g., Schmidt-Ott et al., 1980; Muller et al., 1991) and in order

    to derive upper estimates of the currents that we may expect, we

    have chosen a value of Y 1:0 for our calculations. Finally,

    following Plane (2003) who argues that MSPs should be some sort

    of iron or silicon oxides, we consider refractive indices for Fe2O3and SiO, respectively (see http://www.astro.uni-jena.de/Labora-

    tory/OCDB/oxsul.htm and http://luxpop.com/RefractiveIndexList.

    htm, for corresponding internet data bases).Assuming that only single photoelectrons are emitted, that the

    photoemission/photodetachment processes are independent, and

    finally that the maximum charge is collected by the electrode

    during the duration of the first sampling interval after the flash,

    we estimate the corresponding maximum currents as

    IMSPmax

    Z1rmin

    ZveDt2:5 cm

    Zhc=Wp110 nm

    dNpdrp

    dF

    dl sMSPphotorp;l P drp dl dl

    e

    Dt(2)

    where rmin 0:2 nm is the minimum assumed size of MSPs

    following the original proposal of Hunten et al. (1980) and taking

    into account the argument that immediate re-condensation in ameteor trail is highly unlikely (Plane, 2000). Note that the exact

    value of rmin is not critical for the value of Imax because these

    smallest particles contribute only weakly to the overall current

    since sMSPphoto / r3p. Furthermore, ve is the velocity of a photoelec-

    tron, Dt 10ms is the sampling interval in the fast data channel(and hence also the charge integration time of our integrator

    circuit), h is Plancks constant, cis the speed of light, and Wp is the

    threshold energy for photoionization/photodetachment of a

    particle, i.e., the workfunction or electron affinity of thecorresponding material. dF=dl is the number of photons per

    wavelength interval emitted in one flash and l is the distance from

    the particle detector. P S=4pl2 is the probability that the

    photoelectron is emitted towards the detector electrode with area

    S. dNp=drp is the number density of MSPs per size interval drp, and

    dl and dl are the length and wavelength elements over which the

    integrations above are carried out. Finally, e is the charge of

    an electron. Note that the integration over the wavelengthl starts at 110 nm because of the transmission properties of the

    MgF2-window of the Xe-flashlamp.

    Likewise, photocurrents due to photoionization/photodetach-

    ment of the species in Table 1 can be calculated as

    Imax ZveDt

    2:5 cm

    Zhc=Wi110 nm

    dFdl sil P ni dl dl

    ! eDt

    (3)

    % Fitot si ni

    ZveDt2:5 cm

    P dl

    !e

    Dt(4)

    where ni is the number density of species i, sil is the cross

    section for photoionization/photodetachment of species i at

    wavelength l, and Wi is the corresponding threshold energy

    (i.e., workfunction or electron affinity).

    Note that for approximation (4) we have assumed thatRdF=dlsildl may be approximated as F

    itot si where F

    itot is

    the total number of photons emitted in one flash with an energy

    larger than Wi and si is the average cross section over the

    corresponding wavelength interval.

    Altitude profiles of expected maximum currents based onEqs. (2) and 4 and the altitude profiles shown in Fig. 9 assuming

    either Fe2O3 or SiO particles are presented in Fig. 10. This figure

    clearly shows that photoionization/photodetachment of MSPs

    quantitatively explains the observed photocurrents. In contrast,

    however, potential contributions to the detected photocurrents

    from the photoionization of metal atoms or from the photo-

    detachment of negative ions reach maximum currents ofo1 pA

    and can hence be safely ignored. Even in the case of O21Dg,

    which is by far the most abundant of the considered species, we

    see that expected currents reach a maximum of$100 pA. This is

    still a factor of 10100 less than the currents observed during our

    rocket flight (gray shaded area). Likewise, currents due to NO can

    only dominate the measurements above an altitude of$90 km. In

    the companion paper by Strelnikova et al. (2008) we will comeback to this point and try to estimate the actual NO density above

    ARTICLE IN PRESS

    Table 1

    Electron production by photoionization/photodetachment

    Species Threshold energy (eV) Threshold wavelength (nm) Cross section cm2 Reference

    NO 9.25 134.0 1 1018 Watanabe et al. (1953)

    O21Dg 11.1 111.8 2 10

    18 Clark and Wayne (1970)

    Na 5.14 241.0 1 1019 Bautista et al. (1998)

    Fe 7.87 157.5 2 1018 Kelly and Ron (1971)NO3 3.90 317.9 1 10

    19 Smith et al. (1979)

    CO3 2.90 427.6 1 1018 Cosby et al. (1976)

    O2 0.43 2883.6 1 1018 Cosby et al. (1976)

    Fig. 9. Altitude profiles of the concentration of constituents in the mesosphere/

    lower thermosphere which could potentially be photoionized, or from which an

    electron could be photodetached. Concentrations of sodium and iron atoms (black

    lines) are taken from Plane (2003), of nitric oxide (blue line) from Siskind et al.

    (1998) for 70N and equinox conditions, O2(1Dg) number densities (yellow line)

    are from Gumbel et al. (1998), number densities of negative ions (red lines) from

    Thomas and Bowman (1985), and of meteor smoke particles (green lines) from

    Megner et al. (2006).

    M. Rapp, I. Strelnikova / Journal of Atmospheric and Solar-Terrestrial Physics 71 (2009) 477485 483

    http://www.astro.uni-jena.de/Laboratory/OCDB/oxsul.htmhttp://www.astro.uni-jena.de/Laboratory/OCDB/oxsul.htmhttp://luxpop.com/RefractiveIndexList.htmhttp://luxpop.com/RefractiveIndexList.htmhttp://luxpop.com/RefractiveIndexList.htmhttp://luxpop.com/RefractiveIndexList.htmhttp://www.astro.uni-jena.de/Laboratory/OCDB/oxsul.htmhttp://www.astro.uni-jena.de/Laboratory/OCDB/oxsul.htm
  • 8/3/2019 Eco Ma 1

    8/9

    90 km on the basis of our measured photocurrents. Below 90 km,

    however, measured photocurrents can only be explained byphotoionization/photodetachment of MSPs. Looking at Fig. 10 we

    also find it noteworthy to point out that differences between

    currents owing to photoionization of neutral particles and

    photodetachment of electrons from negatively charged particles

    are minor. The reason for this lies in the fact that the absorption

    cross sections vary as $r3p=l, i.e., the cross sections at the shortest

    relevant wavelengths are considerably larger than the ones at

    longer wavelengths. Hence, it does not matter too much if the

    integration over wavelength starts at corresponding energies

    of 2 or 5.5eV.

    In conclusion of this section, we may state that all available

    arguments do support our initial assumption, i.e., that the

    recorded photocurrents are due to photoelectrons from MSPs, at

    least at altitudes below $90km.

    5. Conclusions

    In the current study we have introduced a new detector design

    for the measurement of MSPs in the middle atmosphere. The new

    detector combines the classical Faraday cup-design of Havnes et

    al. (1996) with a xenon-flashlamp for the active photoionization/

    photodetachment of MSPs. Accordingly, our new detector has two

    measurement channels, one for the detection of a priori charged

    MSPs which can penetrate into the Faraday cup and deposit their

    charge signature on the electrode, and one for the detection of

    very short photoelectron pulses which are actively created by

    xenon-flash photons. The first successful launch of this detectortook place in September 2006 from the Andya Rocket Range and

    yielded high quality data in both data channels. The direct Faraday

    cup measurements showed signatures of negatively charged MSPs

    in the limited altitude range between 80 and 90 km in qualitative

    agreement with earlier measurements with this technique. In our

    companion paper by Strelnikova et al. (2008) we show quantita-

    tively that this limited altitude range is most probably a mere

    consequence of aerodynamical effects and does not reflect a

    layering process in the atmosphere. In agreement with this

    conclusion, measured photocurrents were detected in a much

    broader altitude range between 60 and 110 km. We have then

    devised a laboratory experiment which verifies that measured

    charge pulses truly originate from photoelectrons. With the same

    laboratory setup we could further demonstrate that the observeddisappearance of photocurrents at 60 km on the descent of the

    rocket flight is a direct consequence of the mean free path of

    the photoelectrons in the ambient atmosphere. Interestingly, this

    effect makes our measurement more sensitive if the instrument is

    in the wake of the rocket than when it is in the ram. Finally, we

    investigated the question whether the observed photocurrents

    originated from MSPs or whether they could be due to other

    ionizable species like NO, excited molecular oxygen, metal atoms,

    or negative ions. This analysis shows that all these alternativespecies fall short to explain the observed photocurrents by several

    orders of magnitude below an altitude of 90km. Below an altitude

    of 90km, we succeeded to explain the measured currents

    quantitatively by assuming that they originated from the photo-

    ionization/photodetachment of MSPs where we assumed that the

    particles consisted of iron or silicon oxides and had a photo-

    ionization/photodetachment yield Y 1. Above 90 km, measured

    photocurrents are most likely due to NO and can hence be used to

    derive NO concentrations. This is discussed in detail in the

    companion paper by Strelnikova et al. (2008).

    In summary, our results demonstrate that the active photo-

    ionization and subsequent detection of photoelectrons provides a

    promising new tool for the study of meteor smoke particles in the

    middle atmosphere. Importantly, this new technique does not relyon the a priori charge of the particles, neither is the accessible

    particle size range severely limited by aerodynamical effects.

    In the companion paper by Strelnikova et al. (2008) we apply the

    expressions for the photocurrents presented in this study to

    derive quantitative information on MSP properties which were so

    far unaccessible by any other technique.

    Acknowledgments

    We are indebted to F.-J. Lubken for his support of the ECOMA

    project, to J. Gumbel and M. Friedrich for valuable comments to

    the manuscript, to von Hoerner System GmbH in Schwetzingen,

    Germany, for their excellent contribution to the development of

    the ECOMA flight electronics, to H.-J. Heckl for building thehardware of the ECOMA-detector, and to Misha Khaplanov,

    Department of Meteorology of Stockholm University, for his help

    with measuring the spectrum of the xenon-flashlamp. We also

    appreciate the excellent support by the mobile rocket base of the

    German Space Center. This work was supported by the German

    Space Agency (DLR) under Grant 50 OE 0301 (Project ECOMA).

    References

    Amyx, K., Sternovsky, Z., Knappmiller, S., Robertson, S., Horanyi, M., Gumbel, J.,2008. In-situ measurement of smoke particles in the in the wintertime polarmesosphere between 80 and 85 km altitude. Journal of Atmospheric and Solar-Terrestrial Physics 70, 6170.

    Barjatya, A., Swenson, C.M., 2006. Observations of triboelectric charging effects on

    Langmuir-type probes in dusty plasma. Journal of Geophysical Research 111,A10302.

    Bautista, M.A., Romano, P., Pradhan, A.K., 1998. Resonance-averaged photoioniza-tion cross sections for astrophysical models. Astrophysical Journal SupplementSeries 118, 259265.

    Bohren, C.F., Huffman, D.R., 1983. Absorption and Scattering of Light by SmallParticles. Wiley, New York.

    Clark, I.D., Wayne, R.P., 1970. The absolute cross section for photoionization ofO2

    1Dg. Molecular Physics 18, 523531.

    Cosby, P.C., Ling, J.H., Peterson, J.R., Moseley, J.T., 1976. Photodissociation andphotodetachment of molecular negative ions. III formed in CO2=O2=H2Omixtures. Journal of Chemical Physics 65, 52675274.

    Croskey, C., Mitchell, J., Friedrich, M., Torkar, K., Hoppe, U.-P., Goldberg, R., 2001.Electrical structure of PMSE and NLC regions during the DROPPS program.Geophysical Research Letters 28, 14271430.

    Croskey, C.L., Mitchell, J.D., Friedrich, M., Schmidlin, F.J., Meriwether, J.W., 2003.Initial results of a rocket-based study of gravity wave effects on photoioniza-tion in the middle atmosphere. Advances in Space Research 32 (5), 741746.

    Gelinas, L.J., Lynch, K.A., Kelley, M.C., Collins, S., Baker, S., Zhou, Q., Friedman, J.S.,1998. First observation of meteoritic charged dust in the tropical mesosphere.Geophysical Research Letters 25, 40474050.

    ARTICLE IN PRESS

    Fig. 10. Expected maximum currents to be detected due to photoelectrons from

    photoionization/photodetachment from the species shown in Fig. 9 by the ECOMA-

    particle detector and assuming Fe2O3 or SiO particles. The violet (green) solid lines

    show results assuming photodetachment from Fe2 O3 SiO

    particles for an

    electron affinity of 2 eV (Wang et al., 1996). The corresponding dashed lines show

    corresponding results assuming photoionization of neutral particles with a

    workfunction of 5.5 eV. The gray shaded area in each panel indicates the range

    of currents observed by the ECOMA-particle detector.

    M. Rapp, I. Strelnikova / Journal of Atmospheric and Solar-Terrestrial Physics 71 (2009) 477485484

  • 8/3/2019 Eco Ma 1

    9/9

    Gumbel, J., 2001. Aerodynamic influences on atmospheric in situ measurements

    from sounding rockets. Journal of Geophysical Research 106, 1055310563.Gumbel, J., Murtagh, D.P., Espy, P.J., Witt, G., Schmidlin, F.J., 1998. Odd oxygen

    measurements during the noctilucent cloud 93 rocket campaign. Journal of

    Geophysical Research 103, 2339923414.Havnes, O., Naesheim, L.I., 2007. On the secondary charging effects and structure of

    mesospheric dust particles impacting on rocket probes. Annals de Geophysi-

    que 25, 623637.Havnes, O., Trim, J., Blix, T., Mortensen, W., Nsheim, L.I., Thrane, E., Tnnesen, T.,

    1996. First detection of charged dust particles in the Earths mesosphere.Journal of Geophysical Research 101, 1083910847.

    Hedin, J., Gumbel, J., Rapp, M., 2007. On the efficiency of rocket-borne particle

    detection in the mesosphere. Atmospheric Chemistry and Physics 7,

    37013711.Horanyi, M., Gumbel, J., Witt, G., Robertson, S., 1999. Simulation of rocket-borne

    particle measurements in the mesosphere. Geophysical Research Letters 26,

    15371540.Horanyi, M., Robertson, S., Smiley, B., Gumbel, J., Witt, G., Walch, B., 2000. Rocket-

    borne mesospheric measurement of heavy mb10amu charge carriers.

    Geophysical Research Letters 27, 38253828.Hunten, D.M., Turco, R.P., Toon, O.B., 1980. Smoke and dust particles of meteoric

    origin in the mesosphere and stratosphere. Journal of Atmospheric Science 37,

    13421357.Kelly, H., Ron, A., 1971. Photoionization cross section of the neutral iron atom.

    Physical Review Letters 26, 13591361.Lubken, F.-J., 1999. Thermal structure of the Arctic summer mesosphere. Journal of

    Geophysical Research 104, 91359149.Lynch, K.A., et al., 2005. Multiple sounding rocket observations of charged dust in

    the polar winter mesosphere. Journal of Geophysical Research 110, A03302.Megner, L., Rapp, M., Gumbel, J., 2006. Sensitivity of meteoric smoke distribution

    to microphysical properties and atmospheric conditions. Atmospheric Chem-

    istry and Physics 6, 44154426.Muller, U., Amman, M., Burtscher, H., Schmidt-Ott, A., 1991. Photoemission from

    clean and oxygen-covered ultrafine nickel particles. Physical Review B 44,

    82848287.Plane, J.M.C., 2000. The role of sodium bicarbonate in the nucleation of noctilucent

    clouds. Annals de Geophysique 18, 807814.Plane, J.M.C., 2003. Atmospheric chemistry of meteoric metals. Chemical Reviews

    103, 49634984.

    Rapp, M., Thomas, G.E., 2006. Modeling the microphysics of mesospheric iceparticles: assessment of current capabilities and basic sensitivities. Journal ofAtmospheric and Solar-Terrestrial Physics 68, 715744.

    Rapp, M., Gumbel, J., Lubken, F.-J., 2001. Absolute density measurements in themiddle atmosphere. Annals de Geophysique 19, 571580.

    Rapp, M., Hedin, J., Strelnikova, I., Friedrich, M., Gumbel, J., Lubken, F.-J., 2005.Observations of positively charged nanoparticles in the nighttime polarmesosphere. Geophysical Research Letters 32, L23821.

    Rapp, M., Strelnikova, I., Gumbel, J., 2007. Meteoric smoke particles: evidence from

    rocket and radar techniques. Advances in Space Research 40, 809817.Rosinski, J., Snow, R.H., 1961. Secondary particulate matter from meteor vapors.Journal of Metals 18, 736745.

    Schmidt-Ott, A., Schurtenberger, P., Siegmann, H., 1980. Enormous yield ofphotoelectrons from small particles. Physical Review Letters 45, 12841287.

    Schulte, P., Arnold, F., 1992. Detection of upper atmospheric negatively chargedmicroclusters by a rocket borne mass spectrometer. Geophysical ResearchLetters 19, 22972300.

    Siskind, D.E., Barth, C.A., Russel III, J.M., 1998. A climatology of nitric oxide in themesosphere and thermosphere. Advances in Space Research 21, 13531362.

    Smith, G.P., Lee, L.C., Cosby, P.C., 1979. Photodissociation and photodetachment ofmolecular negative ions. VIII Nitrogen oxides and hydrates. Journal of ChemicalPhysics 71, 44644470.

    Strelnikova, I., Rapp, M., Raizada, S., Sulzer, M., 2007. Meteor smoke particleproperties derived from Arecibo incoherent scatter radar observations.Geophysical Research Letters 34, L15815.

    Strelnikova, I., et al., 2008. Measurements of meteor smoke particles during theECOMA-2006 campaign: 2. results. Journal of Atmospheric and Solar-Terrestrial Physics, this issue, doi:10.1016/j.jastp.2008.07.011.

    Summers, M., Siskind, D.E., 1999. Surface recombination of O and H2 on meteoricdust as a source of mesospheric water vapor. Geophysical Research Letters 26,1837.

    Thomas, L., Bowman, M.R., 1985. Model studies of the D-region negative-ioncomposition during day-time and night-time. Journal of Atmospheric andTerrestrial Physics 47, 547556.

    Voigt, C., et al., 2005. Nitric acid trihydrate (NAT) formation at low NATsupersaturation in polar stratospheric clouds (PSCs). Atmospheric Chemistryand Physics 5, 13711380.

    Wang, L.S., Wu, H., Desai, S.R., 1996. Sequential oxygen atom chemisorption onsurfaces of small iron clusters. Physical Review Letters 76, 48534856.

    Watanabe, K., Marmo, F.F., Inn, E.C.Y., 1953. Photoionization cross section of nitricoxide. Physical Review 91, 11551158.

    ARTICLE IN PRESS

    M. Rapp, I. Strelnikova / Journal of Atmospheric and Solar-Terrestrial Physics 71 (2009) 477485 485

    http://dx.doi.org/10.1016/j.jastp.2008.07.011http://dx.doi.org/10.1016/j.jastp.2008.07.011http://dx.doi.org/10.1016/j.jastp.2008.07.011