Volume 113, Issue D17
Aerosol and Clouds
Free Access

A brief review of the problem of lightning initiation and a hypothesis of initial lightning leader formation

Danyal Petersen

Danyal Petersen

Atmospheric Sciences, Desert Research Institute, Reno, Nevada, USA

Search for more papers by this author
Matthew Bailey

Matthew Bailey

Atmospheric Sciences, Desert Research Institute, Reno, Nevada, USA

Search for more papers by this author
William H. Beasley

William H. Beasley

School of Meteorology, University of Oklahoma, Norman, Oklahoma, USA

Search for more papers by this author
John Hallett

John Hallett

Atmospheric Sciences, Desert Research Institute, Reno, Nevada, USA

Search for more papers by this author
First published: 10 September 2008
Citations: 67

Abstract

[1] A brief review of hypothesized mechanisms of lightning initiation is presented, with the suggestion that these mechanisms provide an incomplete picture of lightning initiation. This is followed by two ideas: (1) a combination of previously hypothesized lightning initiation mechanisms as a means for local intensification of the thundercloud electric field, and (2) a process for the formation of a hot lightning leader channel that is analogous to the space leader phase of the laboratory negative stepped leader. Thundercloud electric field observations have consistently yielded peak values that are an order of magnitude weaker than the dielectric strength of air. Various mechanisms have been proposed to explain how lightning can initiate in such weak electric fields, including hydrometeor-initiated positive streamers and cosmic ray-initiated runaway breakdown. The historically favored positive streamer mechanisms are problematic due to requiring electric fields two to three times larger than peak observed fields. The recently favored runaway breakdown mechanisms appear capable of developing in conditions comparable to peak observed fields although it is not clear how these diffuse discharges can lead to creation of a lightning leader. This paper proposes a solution whereby runaway breakdown and hydrometeor-initiated positive streamer systems serve to locally intensify the electric field. Following this local field intensification, it is hypothesized that formation of the initial lightning leader channel is analogous to the formation of a space leader in a laboratory negative stepped leader.

1. Introduction

[2] One of the more puzzling questions about lightning is the fact that it is somehow initiated in storm clouds in which the observed electric fields are an order of magnitude too weak. Observational evidence has consistently yielded peak thundercloud electric fields that are an order of magnitude weaker than the dielectric strength of air [Marshall et al., 1995]. It may be the case that the strongest electric field regions are compact and have simply eluded sampling. It may also be the case that there exist prelightning breakdown processes that are capable of developing in weaker field conditions and that lead to formation of the lightning leader system.

[3] Loeb [1966] first suggested that positive streamers could be initiated on polarized raindrops in high electric field regions and could develop into branching systems capable of funneling and concentrating the diffuse negative space charge of the thundercloud, leading to the initiation of a negative stepped leader. However, the electrical conductivity of positive streamer tails falls off rapidly in time due to electron attachment and recombination processes [Phelps and Griffiths, 1976] implying that Loeb's funneling mechanism cannot occur. A variation on Loeb's idea was introduced by Phelps [1974], who found that positive streamers developing in strong electric fields undergo intensification whereby they increase their net positive charge while depositing the excess negative charge in their wake. Phelps suggested that this deposition of negative charge by an intensifying positive streamer system could serve as an effective “funneling” mechanism, and that a series of such systems traversing the same volume could lead to significant retrograde movement of negative charge thus creating a strong local intensification of the electric field at the collective streamer system origin. Griffiths and Phelps [1976] developed a simple computer model of this mechanism and found that a series of less than 10 overlapping positive streamer systems, each initiated as a corona on a nearby hydrometeor, may be capable of locally intensifying the electric field by up to an order of magnitude. However, the electric fields required for this mechanism to produce significant electric field intensification appear to be at least a factor of two larger than the strongest observed thundercloud electric fields.

[4] It is possible that there exist compact regions of electric field in thunderclouds that are sufficiently strong and extensive to support the positive streamer system mechanism. Such regions may occur when small parcels of oppositely charged hydrometeors approach each other due to turbulent motions, possibly near the updraft/downdraft interface region where charge separation occurs. It is also possible that such regions may occur due to discharge processes based on the runaway breakdown mechanism. Gurevich et al. [1992] proposed that runaway breakdown could explain the ability of lightning to initiate in the otherwise weak background electric field. It turns out that the electric field strength necessary for a runaway breakdown discharge corresponds approximately to peak observed thundercloud electric fields [Marshall et al., 1995], although distances on the order of a kilometer are required for a significant discharge to develop. Gurevich et al. [1999] hypothesized that a large quantity of seed relativistic electrons, generated in an extensive cosmic ray shower, could lead to a significant runaway breakdown event capable creating a large volume of cool plasma. This plasma, although rapidly immobilized by attachment to neutrals, could be capable of some degree of polarization that could generate strong electric fields at its extremities. It is suggested that the core of the polarized plasma, characterized by the largest ion densities, might structurally resemble the tip of a streamer and therefore be capable of making a direct transition to a streamer discharge and perhaps lightning. Dwyer [2005] suggested that this mechanism may not be viable due to the lateral spreading of a runaway avalanche and corresponding dilution of the resulting plasma, but maintained that continued runaway avalanches in the same volume, sustained by feedback processes, could result in a zone of locally intense electric field near the propagating discharge boundary. It is suggested that this locally intense electric field could attain values in excess of 1 MV·m−1 at sea level pressure and thus support “conventional” breakdown processes.

[5] In this paper two distinct phases of lightning initiation are discussed. The first concerns the mechanism(s) whereby the weak thundercloud electric field may be locally intensified. The hydrometeor-initiated positive streamer system and cosmic ray-initiated runaway breakdown mechanisms are briefly reviewed, after which it is suggested that a serial combination of these processes may provide a better means for local electric field intensification. In this combination scenario it is suggested that a runaway breakdown event first generates a region of locally intensified electric field. This field need only be strong enough to support the generation of positive streamers on nearby hydrometeors and their subsequent development into positive streamer systems. The positive streamer systems result in “funneling” negative charge back toward the streamer system origin, further intensifying the electric field at the origin by an order of magnitude or more. This combination of processes takes advantage of both the lower electric field requirements for runaway breakdown and the local electric field amplification property of a series of overlapping positive streamer systems. It is proposed as a more favorable solution over runaway breakdown or positive streamer systems alone because it relaxes the requirement that runaway breakdown directly support lightning leader formation while offering a means to jumpstart the positive streamer system mechanism of further electric field intensification.

[6] The second phase of lightning initiation that is discussed is the creation of the initial lightning leader system. If we assume that the thundercloud electric field can be locally intensified by processes such as runaway breakdown and/or positive streamer systems (or even that such local electric field regions may already exist by more ordinary means), then we are left with the problem of how the existence of an intense electric field can transition into a hot lightning leader system. Most, if not all, literature dealing with lightning initiation, including those works referenced in this paper, typically explain the final step of leader formation by reference to a presumed “conventional” process. Other than the positive streamer system mechanism of local electric field intensification, we are unaware of the details of a “conventional” lightning initiation process and posit that none more detailed than the positive streamer mechanism has actually been hypothesized. We therefore present as a candidate the idea that the initial lightning leader may form in a manner analogous to the “space” leader element of the laboratory negative stepped leader. While this claim is largely hypothetical, we argue for its plausibility based on known properties of laboratory space leaders and the hypothesized properties of the hydrometeor-initiated positive streamer system.

2. Mechanisms of Locally Intensifying the Thundercloud Electric Field

2.1. Branching Positive Streamer System

[7] The fundamental mechanism of dielectric breakdown in a gas is the electron avalanche. In a gas of number density N subject to an electric field E, free electrons are accelerated, and if E is sufficiently strong, cause a net increase in free electrons by impact ionization of neutral gas particles. A net ionization coefficient α′ can be defined as
equation image
where α and η are functions of E/N and represent production and removal, respectively, of free electrons per unit distance along a path s parallel to E. The number of free electrons Ne changes along s according to
equation image
Integration of equation (2) over the distance Δs (= sfso) parallel to E yields
equation image
For α′ > 0, Ne increases exponentially with distance forming an “avalanche” of electrons. In air α′ = 0 at E/N ≈ 1.24 × 10−19 V·m2 [Bazelyan and Raizer, 1998], yielding a threshold breakdown value or “dielectric strength” of Eb = 3 MV·m−1 for air at STP.

[8] Localized regions of strong E near highly curved electrodes can result in localized discharges called “coronas”. In a corona discharge, free electrons that are accelerated by E in the region where α′ > 0 yield electron avalanches. For the anodic case in air, the avalanching electrons are quickly collected by the anode, leaving behind a relatively immobile positive space charge around the anode that alters the local electric field. If electron avalanches occur uniformly across the anode surface, the resultant positive space charge tends to be smoothly distributed causing an overall reduction of electric field strength near the anode. If, however, a local nonuniformity of positive space charge develops, a localized region of intensified electric field will exist around it (Figure 1). New electron avalanches that develop in this intensified electric field will expose more positive space charge, effectively extending the positive space charge region into the electrode gap. If the background electric field in the electrode gap is strong enough, the positive space charge region can continue to extend into the electrode gap.

Details are in the caption following the image
Formation of a positive streamer in the intensified electric field near a curved anodic surface. Electron avalanches near the anode (where α′ > 0) deposit a region of positive space charge protruding from the anode surface. The electric field is intensified ahead of this space charge region, resulting in more electron avalanches that deposit positive space charge further into the electrode gap while neutralizing the previously generated positive space charge. In a sufficiently strong intergap electric field E, this process can continue unabated, effectively propagating positive space charge across the electrode gap.
[9] An anodic discharge that continues to propagate in the manner described above is called a “positive streamer” [Meek, 1940]. Dawson and Winn [1965] modeled a positive streamer as an electrically isolated volume of positive space charge tailed by a net-neutral plasma tail of negligible conductivity. They estimated the net positive charge and density required for propagation to be around 108 positive ions concentrated in a sphere of 30 μm radius. Phelps and Griffiths [1976] described the streamer plasma tail as initially conductive, becoming rapidly nonconductive due to attachment and recombination processes. Effective conductive lengths of the plasma tail were estimated to be on the order of centimeters. Allen and Mikropoulos [1999] empirically determined the minimum electric field necessary for the stable propagation of a positive streamer in air, or the “stability field”(Est), as
equation image
where h is the absolute humidity in grams per cubic meter and δ is the ratio of the local air density to the density of air at STP. For a standard atmosphere at water saturation and an altitude of 5 km, δ ≈ 0.74 and h ≈ 1.6 g/m3 yielding Est ≈ 330 kV/m. They also empirically determined positive streamer velocities vstr in ambient electric fields Ea as
equation image
where Ea is greater than Est.

[10] Just as electric charge has two polarities, so does the streamer discharge. Cathodic streamer discharges, called negative streamers, are very similar to positive streamers, with the primary difference being the orientation of the electron avalanches relative to the streamer head. In the case of positive streamers the electron avalanches are directed toward the streamer head, while for negative streamers they are directed outward. The electric field required to sustain negative streamers is about twice that of positive streamers due to the self-diffusing nature of the discharge [Bazelyan and Raizer, 1998]. Because of this larger electric field requirement, negative streamers are much less likely to occur than are positive streamers given the same conditions. This asymmetry between positive and negative streamers leads to the favoring of positive streamers as the most likely streamer discharge in weak thundercloud electric fields.

[11] Phelps [1974] studied the behavior of positive streamers when subjected to electric fields in excess of Est, and found that they increase in positive charge and radius with distance and deposit negative charge in their wake equal to the positive charge gained. Phelps constructed a simple analytical model describing a system of intensifying and branching streamers by extending the streamer model of Dawson and Winn [1965] where a streamer is represented as a propagating compact volume of positive space charge (the “tip”) followed by a tail of negligibly conducting plasma. For a single positive streamer propagating in the direction s of the ambient electric field Ea, the streamer energy budget is formulated as
equation image
where q is the net positive charge carried in the streamer tip, u is the potential energy stored in the streamer tip, Ea is the ambient electric field strength, and Est is the electric field strength required for stable propagation of positive streamers (the “stability” field). This model is extended to a system of N intensifying and branching positive streamers, each streamer characterized by a potential energy 〈u〉 and a tip charge 〈q〉, propagating in the direction of E. For simplicity, a static equilibrium is assumed between continuous growth of individual streamers (increasing 〈u〉 and 〈q〉) and streamer branching (decreasing 〈u〉 and 〈q〉), allowing 〈u〉 and 〈q〉 to be formulated as constants. A further simplifying assumption is made that assumes the electrostatic potential energy between the positively charged streamer tips and the deposited negative charge to be negligible in comparison to the potential energy stored in the streamer tips. Under these simplifying assumptions, the total streamer system potential energy can be defined as U = Nu〉 and the total streamer system charge as Q = Nq〉, yielding the final energy budget equation:
equation image
Substituting U = (〈u〉/〈q〉)Q, rearranging and integrating over a distance Δs = sfso yields
equation image
where Qo and Qf are the initial and final charges of the system “front”. It can be seen that when Ea > Est, Q increases exponentially with distance. Since the individual positive streamers are assumed to have a constant charge 〈q〉, the exponential growth of Q is manifest as an exponential increase in the number N of positive streamers comprising the streamer system. More interestingly, charge conservation requires that a quantity of negative charge is deposited in the wake of the system with the total charge equal to the quantity −(QQo).

[12] Griffiths and Phelps [1976] formulated a discrete approximation to this model as a basis for a computer simulation, including a potential energy term to account for the electrostatic interaction of the positively charged streamer tips and the deposited negative space charge. A schematic of their discretized model is shown in Figure 2, showing a conical geometry characterized by a forward stepwise propagation of a growing disk-shaped streamer front and the deposition of negatively charged disks in the passed cone volume. In this computer simulation, an additional feature is added whereby multiple streamer systems are propagated through the same volume in a serial fashion. Each streamer system passage alters the local electric field environment due to the associated charge separation, setting the stage for the next streamer system pass. This feature allows for investigation of the cumulative effect of multiple streamer systems on the overall electric field, especially at the streamer system origin. Results from their model runs yielded two results of interest:

Details are in the caption following the image
Schematic of a simple model of an intensifying positive streamer system. Prior to the nth step, the system is represented by the solid line portions extending out from s = 0 to the thick solid line A (the thin lines represent deposited negative charge and the thick line A represents the propagating positive charge). Following the nth step, the system is represented by the propagation forward of the thick solid line A to the position denoted by the thick dashed line C (hence the removal of A) and the additional deposition of negative charge at the thin dashed line B. (reproduced from Griffiths and Phelps [1976]).

[13] 1. After the passage of a small number of streamer systems, typically less than 10, an order-of-magnitude intensification of the electric field is generated near the streamer system origin over a distance scale of a few meters.

[14] 2. The majority of streamer system intensification occurs over the first few meters of development, along with the associated electric field intensification.

[15] The first result suggests that a sequence of positive streamer systems propagating through the same volume can significantly intensify the electric field over a region spanning a few meters, perhaps to the level of direct dielectric breakdown. The second result suggests that the size scale across which this mechanism operates may be on the order of tens of meters or less, requiring that such regions of high ambient electric field may be quite compact. One of the primary concerns with this mechanism is the fact that the streamer stability field as given by equation (5) appears to be about a factor of two larger than observed peak thundercloud electric fields for a given altitude. It may be the case that the necessary regions of high electric field do exist by ordinary means but, due to being compact, have simply eluded observation. It may also be that such regions are rapidly created by the more exotic discharge mechanism known as runaway breakdown.

2.2. Runaway Breakdown

[16] When a free electron moves through a material medium such as a gas, it undergoes elastic and inelastic collisions that result in an effective frictional force. A graph of this frictional force as a function of electron kinetic energy K is given in Figure 3. Up to electron energies of about 0.1 keV, this frictional force is a rapidly increasing function of electron energy. An electron in this energy regime that is subject to an electric field less than Ec will remain in this regime, continually gaining energy from the electric field and losing it due to collisions. However, between about 0.1 keV and 1 MeV, the frictional force is a decreasing function of electron energy. In this energy regime, if the force on the electron due to the electric field is greater than the frictional force due to collisions, the electron will gain energy and accelerate. Such an electron will continue to gain energy and accelerate as long as it exists in the state where the energy gained from the electric field is larger than the energy lost due to friction. As the electron approaches relativistic energies, the friction once again becomes an increasing function of electron energy and the electron eventually reaches equilibrium where energy gained from the electric field is equal to energy lost due to friction. These electrons are termed “runaway” electrons, and are characterized by relativistic energies on the order of 1 MeV. C.T.R Wilson [1924, 1925] first suggested that energetic electrons in Earth’s atmosphere could become runaways under the influence of a strong thundercloud electric field, and that they could generate additional runaway electrons via ionizing collisions with neutrals. Gurevich et al. [1992] suggested that runaway electrons could precipitate a runaway avalanche and that if seeded in the optimal location in a thundercloud by a highly energetic cosmic ray shower, could result in a significant electrical breakdown of the thundercloud environment. They termed this discharge “runaway breakdown” to set it apart from the ordinary types of electrical breakdown that are characterized by low-energy electron avalanches. In addition, they suggested that runaway breakdown may be a crucial process in the initiation of lightning.

Details are in the caption following the image
Plot of the effective frictional force on a free electron moving through air at STP as a function of electron kinetic energy relative to the air. The solid curve is due to inelastic scattering of the electron with air molecules, while the dashed curve includes bremsstrahlung effects. The solid horizontal line eE represents the accelerating force on an electron subject to an electric field E. Runaway electrons occur when an electron has a kinetic energy K between about 0.1 keV and 103 keV and is subject to a force eE that is greater than F. Ec is the critical electric field strength for which all free electrons will run away, and Eth is the minimum electric field strength capable of sustaining runaway electrons (graph taken from Dwyer [2004]).
[17] The population of energetic (MeV) electrons in a runaway avalanche can be characterized as:
equation image
where L is the length traversed by the runaway avalanche, λ is the characteristic avalanche length, No is the initial quantity of runaway electrons and Nre is the final quantity of runaway electrons. On the basis of computer simulations, Dwyer [2003] has estimated λ as
equation image
and Erb, the minimum electric field necessary to support a runaway avalanche, as
equation image
where δ is the ratio of the local air density to the density of air at STP. The values of Erb given by equation (11) are comparable to the maximum observed thundercloud electric fields [Marshall et al., 1995, 2005], suggesting that runaway breakdown may be an important form of electrical breakdown in thunderclouds. However, the values of λ given by equation (10) are on the order of tens of meters or more for the strongest observed thundercloud electric fields, suggesting that runaway avalanches require distances on the order of a kilometer for a significant increase in runaway electrons.

[18] An important feature of runaway breakdown is the generation of large quantities of low energy (<100 eV) electrons produced in the discharge wake by inelastic collisions of runaways with neutrals. Gurevich et al. [2002] has estimated that the ionization of air by a runaway electron ranges from 30 to 50 ions per centimeter of travel under typical atmospheric conditions. These ions form a nonLTE (local thermodynamic equilibrium) plasma that is characterized by rapid electron attachment with characteristic attachment times on the order of 10−7 s [Gallimberti, 1979]. During this short free-electron lifetime, polarization of the plasma may be induced by the thundercloud electric field as illustrated in Figure 4. Gurevich et al. [1999], and Gurevich and Zybin [2001] have suggested that a large runaway discharge, initiated by a powerful cosmic ray shower (K > 1015 eV), could result in polarization of this plasma sufficient to induce a significant local intensification of the electric field at the plasma extremities. It is further suggested that the core of this plasma could be polarized sufficiently to mimic the structure of a streamer tip, perhaps then developing into a lightning leader. Dwyer [2005] has argued against this possibility, suggesting that the seed electrons produced by a cosmic ray shower should encompass a large lateral extent thereby producing a relatively diffuse runaway avalanche. Dwyer hypothesized another form of runaway breakdown, whereby positron and gamma ray feedback spread and sustain a runaway discharge event until collapsing the large-scale electric field to subrunaway strength. Dwyer's simulations suggest that a linearly compact but significant intensification of the electric field may form at the boundary of the discharge region, perhaps enough to support the “conventional” processes of lightning initiation.

Details are in the caption following the image
Diagram of a proposed mechanism of local electric field intensification by a runaway discharge. In addition to generating new relativistic electrons, a runaway discharge generates numerous thermal electrons. Since the relativistic electron population is maximal near the end of the discharge path, the density of thermal plasma is also maximal. Gurevich et al. [1999] and Gurevich and Zybin [2001] has suggested that rapid polarization of this plasma can generate a strong local intensification of the electric field near its extremities.

2.3. Hybrid Scenario for Locally Intensifying the Thundercloud Electric Field

[19] Based on the outcome of the positive streamer system model of Griffiths and Phelps [1976], it appears that the positive streamer system mechanism is capable of boosting the electric field by up to an order of magnitude over a distance of a few meters. Such a boosting, however, requires an initial background electric field about 2 times larger than observed maximum electric fields. Even more important, a seed streamer must be provided in order to initiate the streamer system. The best candidate for the creation of a seed streamer is a corona streamer from a nearby hydrometeor.

[20] Since both the solid and liquid phases of water are electrically conductive, hydrometeors tend to polarize in electric fields causing enhancement of the electric field around their extremities. Observations have shown that, under the proper conditions, this local electric field enhancement is sufficient to support various corona processes, including positive streamers. Early studies of corona on hydrometeors focused on water drops, because of their presence in regions of known lightning development. When subject to strong electric fields, water drops are observed to deform into elongated shapes, with the elongated ends extending and disrupting into a spray of droplets. Richards and Dawson [1971] investigated corona on water drops of radius >2 mm that were falling through air at 1000 mb pressure while subject to positive vertical electric fields. They reported discharge threshold electric field values for positive corona to be around 950 kV·m−1 for uncharged water drops and 550 kV·m−1 for highly positively charged water drops. Griffiths and Latham [1972] investigated corona on water drops of radius 2.7 mm falling through air at 1000 mb and 500 mb pressure and subject to horizontal and positive vertical electric fields. At an air pressure of 1000 mb, threshold electric field values for positive corona on were around 900 kV·m−1 for positive vertical electric fields and around 630 kV·m−1 for horizontal electric fields. At an air pressure of 500 mb, field values changed to 550 kV·m−1 for positive vertical electric fields and 690 kV·m−1 for horizontal electric fields. Crabb and Latham [1974] investigated corona on pairs of colliding water drops in air pressure of 1000 mb and subject to electric fields, using drops of radius 2.7 mm and 0.65 mm that traveled toward each other at a velocity of 5.8 m·s−1. The collisions were often observed to produce temporary elongations of the interacting drop pairs in the form of water filaments up to 20mm in length, with the longer water filaments resulting from more glancing collisions. Threshold electric field values for positive corona ranged from about 500 kV·m−1 for water filaments of length 10 mm, down to about 200 kV·m−1 for water filaments of length 20 mm. The measurements of Crabb et al. represent the lowest corona threshold values for known individual and interacting hydrometeors. Coronas from ice hydrometeors have been less extensively studied, with the most important work done by Griffiths and Latham [1974]. In their experiment, they studied various habits of ice hydrometeors with lengths ranging from 4 to 25 millimeters in length. They found that the electric field required for positive streamer emission for vapor-grown ice crystals ranged from 600–800 kV·m−1 at sea level pressures down to about 400 kV·m−1 at 500 mb. Hailstones provided only slightly lower electric field requirements, with a 25 mm long hailstone yielding threshold fields of 500 kV·m−1 at 1000 mb down to about 360 kV·m−1 at 500 mb. In one of their oft-quoted findings, they found that continuous corona currents could be induced on ice crystals only at temperatures above −18°C. This finding is often taken as evidence that ice crystals cannot contribute to streamer production at colder temperatures such as occur in the upper regions of thunderclouds. However, if we consider that a positive streamer system requires only a single seed streamer and that less than 10 systems may create an order-of-magnitude increase in the local electric field, it may only be necessary for an ice crystal to produce a small number of positive streamers to be viable for lightning initiation. A more recent study by Petersen et al. [2006] showed that individual positive streamers can indeed be generated from ice crystals at temperatures as low as −38°C, in contrast to the misunderstood result of Griffiths and Latham.

[21] Based on the available studies, it appears that the electric fields required for positive streamer generation on hydrometeors exceeds the electric fields required for the positive streamer mechanism of local electric field intensification hypothesized by Griffiths and Phelps [1976]. If we assume that a small population of hydrometeors is capable of going into corona at around 400–500 kV·m−1, then it follows that the emerging seed streamer will immediately develop into an intensifying positive streamer system and possibly lead to a local intensification of the electric field.

[22] As is well known, observations of thundercloud electric fields suggest that no such strong electric fields exist in thunderclouds. It could well be that such regions do exist by ordinary means, and that they have simply eluded observation due to being quite compact. Alternatively, such regions may be formed by action of the runaway breakdown mechanism. Evidence in support of the runaway breakdown hypothesis includes, among other things, a close match between runaway breakdown electric fields and maximum observed thundercloud electric fields. Indeed, runaway breakdown may be one of many viable mechanisms for locally boosting the thundercloud electric fields.

[23] We propose a hybrid mechanism of local thundercloud electric field intensification that takes advantage of both runaway breakdown and hydrometeor-initiated positive streamer systems. This mechanism could equally well conform to the notion of preexisting yet undetected local pockets of strong electric field, with such regions replacing the role of runaway breakdown in the sequence. One scenario begins with an extensive cosmic ray shower seeding a runaway breakdown event that proceeds to generate a region of cool plasma in the high electric field region of the thundercloud. Before attachment renders the plasma nonconductive, the plasma polarizes and creates a local intensification of the electric field that reaches or exceeds the value required for coronas on nearby hydrometeors. A nearby hydrometeor then goes into corona, generating a single positive streamer. This streamer rapidly develop into a positive streamer system, quickly filling the newly formed high-field region of the polarized plasma and further intensifying the electric field near the streamer system origin (Figure 5). The positive feedback on the electric field at the streamer system origin leads to a continued succession of positive streamer systems, rapidly boosting the electric field near the streamer system origin. With the propagation velocity of positive streamers being around 105 m·s−1 and the length scale of interest being around 10 m, the timescale of this process would be on the order of a millisecond.

Details are in the caption following the image
Diagram of the positive streamer mechanism of local electric field intensification. A positive streamer is generated on a hydrometeor in the intensified electric field near runaway-generated polarized plasma, and develops further into a branching positive streamer system. This creates a further intensification of the electric field near the streamer system origin that may initiate another positive streamer system. A series of positive streamer systems may result in an order-of-magnitude intensification of the local electric field.

[24] Another scenario involving runaway breakdown involves the mechanism of Dwyer [2005]. In this case, when the background thundercloud electric field exceeds the runaway threshold, positive feedback on the runaway breakdown creates a forward-propagating region of intensified electric field at the runaway breakdown discharge front. Dwyer's simulation suggests that the electric field in this front can exceed 430 kV·m−1 at around 7000 m above sea level, corresponding to around 1 MV·m−1 at sea level pressure and about 500 kV·m−1 at 500 mb. An important issue with this scenario is the speed with which the discharge front propagates. It is suggested that a well-developed discharge front may attain speeds up to about 106 m·s−1, with the region of electric field exceeding 500 kV·m−1 at 500 mb having a linear extent of around 100 m. The local residence time in this field would be about 0.1 milliseconds, somewhat smaller in magnitude than the estimated time for development of hydrometeor-initiated positive streamer systems. However, it may be possible that this short duration is sufficient to support a brief burst of positive streamer activity that is itself sufficient to support further positive streamer system development and associated local electric field intensification well after the runaway discharge front has passed.

3. Leader-Type Discharges, and a Possible Connection to Lightning Initiation

3.1. Positive Leader

[25] Various studies have characterized the initiation of positive and negative leaders from conducting electrodes [Gallimberti, 1979] as well as initiation of leaders from long (>1 m) conductors floating in electrode gaps [Castellani et al., 1998; Lalande et al., 2002]. Leaders, both in the laboratory and in nature (lightning), consist primarily of a channel of air that is electrically conductive due to thermally driven electron detachment and ionization processes. Laboratory investigations have indicated a minimum temperature for sustained electrical conductivity to be in the range of 1000–2000 K [Aleksandrov et al., 2001; Gallimberti et al., 2002]. The issue of importance in the formation of a leader is the means whereby the leader channel is heated and propagated forward. Figure 6 illustrates the anatomy of the active region of a typical positive laboratory leader. This region consists of the terminal end of the highly conductive leader channel, the growing tip of this channel, and the surrounding streamer zone. Because of the high conductivity of the conductive leader channel, it acts as an anode and generates a very intense electric field ahead of the tip. This intense electric field exceeds the dielectric strength of air near the tip, resulting in the formation of positive streamers that rapidly propagate away from the tip over a distance of a few meters. As the streamers exit the tip, they produce a concentrated current in and around the tip that further heats the tip and propagates it forward. The quantity of streamers necessary for continued positive leader development is constrained to a range of intermediate values. If a leader does not produce enough streamers, it cannot adequately heat the leader tip and channel and thus cannot extend. If the leader produces too many streamers it becomes shrouded in a field-choking positive space charge. This constraint can be quantified as a streamer charge production per unit length of leader development, or Q/L, with laboratory results indicating it to be on the order of 50 μC·m−1 [Gallimberti et al., 2002]. The electric field required for stable positive leader propagation at sea level air pressure varies from 100–200 kV·m−1 for laboratory leaders down to about 10–50 kV·m−1 for lightning leaders, and the propagation rate of positive leaders varies from 104 m·s−1 for laboratory leaders up to 3 × 105 m·s−1 for lightning leaders [Lalande et al., 2002].

Details are in the caption following the image
Schematic of the active region of a propagating positive leader. The conductive leader channel acts as an anode, generating a strong electric field in the vicinity of the tip. Copious amounts of positive streamers are produced at the tip, with the concentrated streamer electron current heating the air just ahead of the leader tip and extending it forward.

3.2. Negative Leader

[26] The active region of a negative laboratory leader has the same general anatomy as a positive laboratory leader, consisting of a leader channel, tip and streamer zone. There is, however, a significant addition to the negative laboratory leader extension process, illustrated in Figure 7. In the negative laboratory leader case, an initial burst of negative streamers is emitted from the main leader tip and propagates through the strong local electric field for a few meters. As illustrated on Figure 7a, this results in the creation of small heated and positively charged stems along the negative streamer paths corresponding to regions of vigorous streamer intensification and/or branching [Reess et al., 1995]. As illustrated on Figure 7b, the positive space charge in these regions leads to the rapid formation of a retrograde positive streamer that propagates back to the main negative leader tip. If the positive streamer emission is energetic enough, a compensatory quantity of negative charge is deposited back into the space stem, resulting in a forward-propagating negative streamer. This discharge sequence is called a “pilot”, and often occurs in a repetitive forward-propagating series with a repetition period of tens of nanoseconds and a propagation velocity on the order of 105 m/s. Pilots serve as a continuous source of retrograde positive streamers that feed into the main negative leader tip. This coupling of pilots to the main negative leader tip provides the majority of the electric current that heats and extends the main negative leader tip.

Details are in the caption following the image
Schematics of the active region of a propagating negative leader. (A) Negative streamers are generated in the strong electric field near the main negative leader tip in much the same way as positive streamer generation in front of a positive leader tip. Due to vigorous streamer development, various regions along the negative streamer tracks are heated and retain excess positive space charge. These regions are called space stems (B) A series of pilot discharges are generated at the space stems, resulting in both retrograde positive streamers that feed into the main negative leader tip and new forward-propagating negative streamers. (C) A space leader forms when a downstream pilot feeds positive streamers to an upstream pilot that, in turn, feeds positive streamers to the main negative leader tip. The space leader is heated by the positive streamer current until it becomes electrically conductive, causing it to extend linearly in a manner consistent with bipolar leader growth. (D) The space leader continues to extend in both directions until its anodic end attaches to the main negative leader channel. This attachment is accompanied by a surge of current and rapid equalization of potential along the space leader channel, resulting in an effective stepwise forward motion of the main negative leader channel.

[27] The situation can arise whereby two or more pilots may exist simultaneously in a linear series in front of the main negative leader tip. This situation is of great interest because it creates the condition whereby one of the pilots may transition into a hot bipolar leader segment [Ortega et al., 1994]. Figure 7c illustrates a linear series of two pilots, with the outer pilot generating positive streamers that are fed into the inner pilot. By being fed current from an upstream pilot and in turning feeding that current downstream to the main negative leader channel, the inner pilot stem is continuously heated via joule heating. After reaching the critical temperature for sustaining electrical conductivity, the inner pilot stem begins to lengthen at both ends in much the same way as the tip of the main negative leader. This elongating conductive structure is called a “space leader”, and is essentially a floating bipolar leader. As illustrated in Figure 7d, the lengthening space leader eventually attaches to the main negative leader channel. At this moment, the space leader is rapidly brought to the potential of the main negative leader tip. This is accompanied by a surge of current and luminosity, and is commonly referred to as a “step”.

[28] Typical Q/L values for negative laboratory leaders are around 100 μC·m−1 [Gallimberti et al., 2002]. Ambient electric fields of 200–300 kV·m−1 are required for propagation at sea level air pressure, nearly double that of positive leaders [Lalande et al., 2002]. The elongation of negative laboratory leaders typically occurs in step lengths on the order of 1 m, with a stepping period of 10–20 μs and velocity of 1–5 × 105 m·s−1. In comparison, negative lightning leaders have step lengths that can exceed several tens of meters and velocities that approach 106 m·s−1. While it may superficially appear that the mechanisms of lightning and laboratory stepped leader extension are identical, debate still exists due the more extreme nature of lightning and the apparent lack of sufficiently detailed observational evidence.

3.3. Hypothesis for Lightning Leader Initiation

[29] While most laboratory studies provide insight into leader initiation on electrodes and large floating conductors, lightning leader initiation usually occurs in regions devoid of such structures and is thus not trivially comparable. Instead of attempting to explain lightning leader formation in terms of electrode-initiated leaders, it may instead be more accurate to consider a comparison to the laboratory space leader as the laboratory space leader shares with lightning the constraint of being initiated in regions devoid of conducting electrodes.

[30] We suggest that the local conditions required for the onset of the pilot process and subsequent formation of a laboratory space leader may also exist in a thundercloud environment. Specifically, we suggest that these conditions may arise in the strong electric field regions that are hypothesized to result from the positive streamer system mechanism of Griffiths and Phelps. Figure 8 illustrates the process as conceived, based on hypothesized properties of positive streamer systems and know properties of the laboratory space leader. According to Griffiths and Phelps [1976], a series of overlapping positive streamer systems may boost the electric field to levels exceeding 1 MV·m−1 over the distance of a few meters. In such conditions, hydrometeors likely undergo both positive and negative corona discharges. The positive streamer discharges would continue to reinforce the positive streamer system effect of locally intensified electric field, while the negative streamers might create space stems in the same way as in the case of the laboratory negative leader (Figure 8b). These space stems could undergo a series of pilot discharges, with the positive streamers continuing on into positive streamer systems (supplanting the hydrometeors as initiators of seed positive streamers). If a small number of pilots becomes serially connected via positive steamers (Figure 8c), then the downstream pilot stems could undergo heating until becoming electrically conductive. These electrically conductive stems could then be further heated by continued pilot-generated streamer activity, resulting in their extension into space leaders. Further growth and extension of a number of serially connected space leaders could lead to their attachment, creating a substantially longer space leader segment (Figure 8d). This attachment process would be very much analogous to the stepping process of the laboratory negative leader, with the role of the main negative leader channel as a positive charge drain being taken by the continuing positive streamer generation on the anodic end. As the space leader channel continues to lengthen, the potential gradient along the conductive leader channel would continue to drop due to continued joule heating, further intensifying the electric field at the leader channel extremities. Given that the local electric field at the leader extremities can sustain continued development, the space leader should continue to lengthen until emerging as a lightning leader channel.

Details are in the caption following the image
Hypothetical schematic of initial lightning leader formation. (A) In locally intense electric field region of thundercloud, corona on hydrometeor generates seed positive streamer. Positive streamer develops into intensifying and branching positive streamer system, further intensifying the electric field near the streamer system origin. (B) Local intensity of electric field now sufficient to generate both positive and negative streamers on nearby hydrometeors. Positive streamers develop into intensifying and branching positive streamer systems, while negative streamers develop less but create space stems. (C) Pilot discharges initiate on space stems, with series connection of space stems allowing inner stems to heat and develop into space leaders. Streamer activity continues at extremities of space stem system. (D) Continued heating and elongation of space leaders allows for rapid elongation and attachment into larger space leader system. Reduction of potential gradient inside the space leader channel further intensifies the electric field at the space leader extremities. Resulting streamer generation at the extremities leads to the creation of new space stems at the cathodic end.

[31] In order for this hypothesized lightning initiation sequence to succeed, a number of constraints must be accommodated. First, the electric field strength required for space stem formation must be met. Observations of laboratory pilot formation have shown that pilots can form in the outer regions of the streamer zone of a laboratory negative leader [Ortega et al., 1994; Reess et al., 1995]. It can be assumed that the electric field in these regions is comparable to the minimum field required for negative streamer propagation, around 750 kV·m−1 at sea level pressure [Gallimberti et al., 2002]. This value, adjusted to 500 mb pressure, is less than the field strength hypothesized to occur near the positive streamer system mechanism. Additionally, the streamers created by the pilots should not develop a local space charge shield that reduces the local electric field and quenches the discharge. In the case of a laboratory negative leader, this constraint is fulfilled via removal of positive charge through the highly conductive leader channel. However, in the case of lightning initiation, no such conductive channel exists. Instead, positive charge removal must be carried out by the propagation of positive streamers away from the pilot. The positive streamer system mechanism is hypothesized to propagate positive space charge to distances on the order of 10 m or more, in addition to further intensifying the electric field at the system origin due to positive streamer intensification. This should, in effect, fulfill the role of positive space charge removal. However, unlike the case of a laboratory negative leader, the embryonic lightning leader must eventually develop its positive end in the general direction of this large deposit of positive space charge. If the accumulated space charge is too great, it may interrupt the positive leader development which would end the discharge. One possibility is that, as the embryonic lightning leader extends out toward the deposited positive space charge, the reduced potential gradient inside the lengthening leader channel results in intensification of the electric field at the positive leader tip sufficient to overcome the shielding effect. Another possibility is that the embryonic lightning leader simply develops around, and thus avoids, the deposited positive space charge. Finally, the emerging embryonic lightning leader must be capable of continued development under the weaker background thundercloud electric field. This constraint depends primarily on the polarization of the leader channel upon emergence into the weaker background field, and is a function of the leader channel length and conductivity. If the leader channel is able to emerge under these constraints, it may be properly stated that lightning has initiated.

4. Summary

[32] For the last few decades, the majority of literature dealing with the problem of lightning initiation has focused primarily on two discharge processes: positive streamers generated from polarized hydrometeors, and relativistic electron avalanches initiated by cosmic rays. These discharge processes may be capable of occurring in the relatively weak thundercloud electric fields, but are unlikely to transform directly into a lightning leader system. Instead, they appear to offer a means whereby the thundercloud electric field may be locally intensified. Typically, the transition from these mechanisms to the onset of the initial lightning leader is described by general reference to “conventional” breakdown processes which are not clearly defined. Thus in our estimation, the problem of onset of the initial lightning leader remains unresolved.

[33] Section 2 includes a brief review of the positive streamer system and the runaway breakdown mechanisms of local electric field intensification, and points out some of the strong and weak points of these mechanisms with respect to lightning initiation. This is followed by the hypothesis that a serial combination of such mechanisms may be more capable of producing a strong local intensification of the electric field in the thundercloud environment. It is suggested that runaway breakdown need not lead directly to the formation of a thermalized leader channel, but instead need only generate a moderate local intensification of the thundercloud electric field necessary to support positive streamer discharges from nearby hydrometeors. It is also mentioned that this moderate intensification may also result from ordinary action of preexisting thundercloud space charge, with the lack of observational evidence of related compact regions of strong electric field being a consequence of insufficient observational sampling. It is then suggested that if these mechanisms moderately locally intensify the thundercloud electric field, the hydrometeor-initiated positive streamer system mechanism may be capable of further locally intensifying the electric field to levels approaching the dielectric strength of air.

[34] Section 3 includes a brief review of some basic leader discharge processes, with focus on the pilot and space leader processes of laboratory negative leader extension. This is followed by a hypothesis that initial lightning leader formation may be functionally similar to the formation of a laboratory pilot and space leader sequence. In place of the main negative leader, it is suggested that hydrometeor-initiated positive streamer systems may create a similar effect by generating the necessary strong local electric field and carrying away the excess positive charge. In addition, given a sufficiently strong local electric field resulting from positive streamer systems, it is suggested that negative streamers may also be generated as coronas on nearby hydrometeors thus allowing for the creation of the necessary “space stem” regions that are required for the formation of pilot discharges.

[35] One possible means of testing this hypothesis would be to investigate the electromagnetic radiation emitted during the initial phase of lightning initiation. We suggest that this investigation look for evidence of the following proposed processes: (1) Moderate local intensification of the electric field, perhaps via a runaway breakdown event. If this step involves runaway breakdown, then the initial radiation should include a strong burst of brehmsstrahlung radiation resulting from runaway electrons colliding with atomic nuclei, and an associated pulse of broadband RF radiation resulting from the large mass of accelerating charges. (2) A series of hydrometeor-initiated positive streamer systems. As individual positive streamers are low-current discharges with conductive lengths on the order of a few centimeters, they likely do not emit significant quantities of RF radiation at wavelengths typically associated with lightning RF emissions. However, during the more vigorous initial portion of streamer system development, the combined electrical currents may be more substantial and thus may emit measurable RF radiation at frequencies above 1 GHz. (3) Formation of space stems and pilots, and the development of pilots into space leaders. Any RF radiation during this phase would likely resemble the previously described radiation as the discharge processes involved are primarily due to coronas and streamers. (4) Attachment of a linear array of space leaders, creating an elongated lightning leader embryo. This stage should be characterized by a series of longer-wavelength RF pulses due to current surges associated with attachment of the space leader segments to each other. As the space leaders are hypothesized to be relatively small compared to step lengths of a typical lightning negative leader, and the corresponding current surges across the channels relatively weak, the characteristic emission pulses should be much weaker than typical stepped leader pulses and may lack the longer wavelength emissions. However, as the embryonic lightning leader system continues to develop, the RF pulses should gradually become more powerful until resembling typical lightning stepped leader pulses.

[36] Further work will involve more accurate investigation of the “conventional” processes described in this paper. Work has begun on further developing the Griffiths and Phelps [1976] model of an intensifying and branching positive streamer system in order to more accurately assess the capacity for such systems to locally intensify the thundercloud electric field. In addition, it is of interest to further develop the pilot/space leader hypothesis of lightning leader formation and to incorporate it into an extended streamer system model as a more complete picture of lightning initiation. Finally, observations will be made of the fine structure of RF emissions during the initial phase of lightning formation with focus on frequencies above 1 GHz.

Notation

  • N
  • - Number density of a gas, m−3
  • E
  • - Electric field, V/m
  • α
  • - Net ionization coefficient, m−1
  • α
  • - Ion production coefficient, m−1
  • η
  • - Ion removal coefficient, m−1
  • s
  • - Path of integration, m
  • Ne
  • - Number of free electrons
  • Eb
  • - Dielectric strength of air
  • Est
  • - Stability field of propagating positive streamers
  • h
  • - Absolute humidity, g/m3
  • δ
  • - Ratio of the local air density to the density of air at STP
  • vstr
  • - Velocity of a positive streamer, m/s
  • Ea
  • - Ambient electric field
  • q
  • - Net positive charge stored in a positive streamer tip
  • q
  • - Average value for q in a positive streamer system
  • Q
  • - Net positive charge stored in a positive streamer system
  • u
  • - Potential energy stored in a positive streamer tip
  • u
  • - Average value of u in a positive streamer system
  • U
  • - Potential energy stored in a positive streamer system
  • K
  • - Electron kinetic energy, eV
  • Kth
  • - Electron kinetic energy threshold for runaway breakdown, keV
  • Ec
  • - Critical electric field strength for which all free electrons will run away
  • Eth
  • - Minimum electric field needed to produce runaway electrons
  • L
  • - Length traversed by a runaway avalanche
  • λ
  • - Characteristic runaway electron avalanche length
  • No
  • - Initial quantity of runaway electrons
  • Nre
  • - Final quantity of runaway electrons
  • Erb
  • - Threshold electric field for a runaway avalanche
  • Glossary

  • Ambient electric field
  • Background thundercloud electric field.
  • Electron attachment
  • Process whereby free electrons become attached to neutral molecules, forming negative ions.
  • Electron recombination
  • Process whereby electrons are recombined with positive ions.
  • Bipolar leader
  • Leader system containing simultaneously developing positive and negative ends.
  • Conventional discharge
  • Reference to an electrical discharge based on the Townsend electron avalanche mechanism.
  • Cool plasma
  • Plasma whose temperature is such that it cannot maintain appreciable electrical conductivity due to rapid electron attachment and recombination processes.
  • Cosmic ray
  • High energy particle of extraterrestrial origin.
  • Cosmic ray shower
  • Shower of elementary particles and high-energy photons that results from collision of a cosmic ray nuclei with the nuclei of an air molecule.
  • Embryonic lightning leader
  • Initial stage of a lightning leader characterized by short length and large electrical resistance.
  • Hydrometeor
  • General term for a water-based particle suspended in Earth’s atmosphere, such as a cloud droplet, raindrop or ice crystal.
  • Leader discharge
  • A specific type of electrical discharge in a gas, characterized by an extending filament of hot, highly electrically conductive gas that is preceded at its leading tip by copious coronal discharge activity.
  • Leader channel
  • The hot, highly electrically conductive filamentary channel of a leader discharge.
  • Leader tip
  • The terminal end of the leader channel where coronal discharges are generated.
  • Lightning leader
  • A leader discharge that forms in or around a thundercloud.
  • Long spark
  • A leader discharge.
  • LTE
  • Acronym for “local thermodynamic equilibrium”. Refers to a plasma in which all particles such as ions and free electrons are in thermodynamic equilibrium.
  • Negative streamer
  • Corona-type discharge characterized by effective advancement of a negative space charge.
  • Negative leader
  • Leader system that propagates negative electric potential.
  • Negative stepped leader
  • Distinct from a typical positive leader due to an additional extension process called 'stepping'.
  • Pilot
  • A series of positive and negative streamers initiated on a compact region of space charge that is subjected to a strong ambient electric field.
  • Plasma
  • A state of matter characterized by a gaseous mixture of electrically charged particles.
  • Positive streamer
  • Corona-type discharge characterized by effective advancement of a positive space charge.
  • Positive leader
  • Leader system that propagates positive electric potential.
  • Relativistic electron avalanche
  • An electron avalanche mechanism that is distinct from the Townsend electron avalanche due to the extremely high energy of the free electrons. This avalanche can develop in ambient electric fields that are 1/10th as strong as those required for the Townsend electron avalanche. Requires an initial supply of relativistic electrons and large propagation distances for significant development.
  • Runaway breakdown
  • Electrical discharge process based on the relativistic electron avalanche.
  • Runaway electron
  • Free electron with a large kinetic energy (≫100 eV), accelerated by an ambient electric field such that energy losses to the environment are a decreasing function of kinetic energy and are less than energy gains from the electric field.
  • Space charge
  • A collection of electrical charges that are distributed in a volume of space.
  • Space leader
  • A bipolar leader that typically forms in the streamer zone of a negative stepped leader and whose attachment to the main negative leader channel results in a ‘stepwise’ extension of the main negative leader channel.
  • Space stem
  • A compact region of space charge in the streamer zone of a negative leader that serves as a starting point for a pilot.
  • Stability field
  • Threshold electric field for stable propagation of a streamer.
  • Stepping
  • Primary process involved in extension of a negative leader, characterized by discrete events of forward propagation of the leader tip.
  • STP
  • Acronym for “standard temperature and pressure”. Defined as 1 atmosphere of pressure and 0° C.
  • Streamer tip
  • Active zone of a streamer, where Townsend electron avalanches occur. Propagates forward as an ionization wave.
  • Streamer tail
  • Cylinder of cool plasma behind the streamer tip, characterized by rapid loss of electrical conductivity due to electron attachment and recombination.
  • Streamer zone
  • Region ahead of a leader tip characterized by streamer activity.
  • Thermal electron
  • Free electron with kinetic energy typically less than a few tens of eV in air. Distinct from a relativistic electron whose energy is typically much greater than 100 eV.
  • Thundercloud
  • Meteorological term for a cloud characterized by lightning discharges.
  • Townsend electron avalanche
  • When a strong ambient electric field is present, free electrons are accelerated to energies in excess of a few eV allowing for impact ionization of neutral atoms and molecules. When the average number of electron impact ionization events exceeds the average number of electron attachment and electron recombination events, the number of free electrons increases exponentially.
  • Acknowledgments

    [37] This work was supported by NSF-EPSCoR grant EPS-0082725 through the Physical Meteorology Program, National Science Foundation, Arlington, VA, USA.