When analyzing the signals from Langmuir probes by numerical simulations, for instance, the currents are usually interpreted by counting the number of charged particles arriving at the probe surface per unit time. Under stationary homogeneous conditions, this model will give the correct average result, but the individual particle contributions are often interpreted incorrectly. As demonstrated by a specific physically realistic example, the contribution from induced charges will be missing. Collected and passing particles will both contribute to the power spectrum of the probe noise. This noise will set the lower limit to the deterministic signals being detectable. Particles that give the current contribute at the same time also to the noise. The emphasis is here on biased probes, but the basic analysis applies to floating probes as well, or conducting macroscopic particles embedded in plasmas. It is argued that a current can be induced into a probe without any charges reaching its surface.

Langmuir probes1–7 are versatile diagnostic tools for studies of low8 and medium temperature plasmas.9 Several numerical simulations have been carried out in order to understand the functioning of such probes.10–15 A probe characteristic is generally obtained by counting the number of charged particles that cross the probe surface per unit time. Although this procedure gives the correct result after averaging, it is argued in the present study that this method can lead to incorrect conclusions when results are compared to the physical performance of these probes, the probe noise in particular. In this simple model, a charge e moving with velocity vector w directed at an angle θ with respect to the normal to a small area element dA contributes with ewcosθδ(rwt), see Fig. 1. For this velocity component, the average charge per unit time arriving at dA is I=e|w|ncosθdA with n being the density of these particles. For particles with an isotropic velocity distribution f(w), the net result becomes I=edA0π/20|w|f(w)cosθdwdσ/4π, where dσ/4π is the fraction of particles that are directed toward dA, crossing a narrow belt around the normal vector16 (see Fig. 1), where dσ=2πdθsinθ. With a Maxwellian velocity distribution for charged particles with mass m and temperature T, the result is I=endAT/2πm. Plasma perturbations such as ion acoustic waves are manifested by time variations of the density, n=n(t).

FIG. 1.

Illustration of the local coordinate system on a conducting sphere, where dσ=2πsinθdθ is the area of the narrow sphere segment.

FIG. 1.

Illustration of the local coordinate system on a conducting sphere, where dσ=2πsinθdθ is the area of the narrow sphere segment.

Close modal

When individual particle current contributions are modeled by δ-pulses, the associated temporal correlation function will also become a δ-function. The associated noise becomes a standard “shot noise” with a flat or “white” power spectrum,17 irrespective of the angle of impact. Only charges arriving at the probe surface will contribute, while passing charges are irrelevant in this standard model. These comments apply solely to the probe–plasma interaction. The observed noise characteristic can be strongly modified by the detecting circuit.4 A characteristic frequency Ω related to an inflow of independent charged particles with mass m taken from a distribution with density n can be constructed by a typical velocity T/m and a length scale given by the average particles separation 1/n1/3, giving ΩNp1/3ωp in terms of a plasma frequency ωp and the plasma parameter NpλD3n, where λD is the Debye length. No meaningful signal in the probe current can be detected at frequencies larger than Ω. Note that Np103 is usually considered large, but then Np1/310 gives a moderate Ω-value.

The present study illustrates the problem by a simple, yet physically realistic and relevant model considering the induced charges in an ideally conducting sphere. Infinitely long cylindrical probes can also be analyzed, but these results will add little to the discussion. The noise contributions from the charged particles differ from the usual “shot noise,” and results are found for the correlation functions and power spectra. It is demonstrated that both collected and passing particles contribute to this noise level, which sets the lower limit to signals that can be identified by such probes.

The model outlined before fails by ignoring the induced charge caused by an approaching particle before it reaches the conducting object. In fact, it does not even need to reach the surface. The following analysis illustrates this process using the method of images. This is a tool explicitly assuming electrostatic conditions by relying on a uniqueness theorem for static electric fields with given boundary conditions.18,19 Strictly speaking, it fails if charges are in motion, since then the electric field need no longer be determined by the boundary conditions alone. For small charge velocities and charge accelerations, where electromagnetic radiation can be ignored, the method is applicable as an approximation similar to Kirchhoff's laws in electronics.20 These latter laws can be understood as corollaries of Maxwell's equations in the low-frequency limit. They are accurate for direct current (DC) circuits, and for alternating current (AC) electronics at frequencies where the wavelengths of electromagnetic radiation are very large compared to the circuits. This limit is also assumed in the following. The arguments are made more detailed in the  Appendix.

Consider an ideally conducting spherical surface as illustrated in Fig. 2. Two point particles, a) and b), with charge e moving with velocity w are shown as well. Particle b) represents one crossing dσ in the direction of dA in Fig. 1. The sphere can be electrically floating or be biased through a thin conducting wire. The first part of the following discussion assumes the latter case, where the reference potential, or “ground,” is taken to be the plasma potential when many charges constituting a plasma surrounding the probe. For this case, the sphere will not disturb the orbits of approaching charges and their velocity vectors will be assumed to be constant. The reference particle charge can have either sign. For an illustration, it is an advantage to use a vertical “cut” in the figure as given in Fig. 3. The moving reference charged particle induces a time varying charge distribution on the surface of the conducting sphere.18,19,21 This charge is mediated by a current i(t) through a connecting wire also shown in Fig. 3. The solution of the problem in terms of image charges is found in many textbooks on electrodynamics,18,19 and only the relevant results need to be reproduced here. A summary figure is included in Fig. 4 with explanations in the caption. Results for many non-interacting charges can be found by superposition. The total net charge (unevenly) distributed over the conducting sphere is q=eR/r=eR/p2+x2 with symbols explained in Fig. 3, see also Fig. 4.

FIG. 2.

A simple physical model illustration of a spherical conducting surface with radius R and moving charged point particles with different impact parameters p. Two selected particles, one passing and another one reaching the spherical surface, are labeled (a) and (b). The angle between the particle trajectory and the connecting wire is not relevant for the analysis, and here, they are shown to be parallel.

FIG. 2.

A simple physical model illustration of a spherical conducting surface with radius R and moving charged point particles with different impact parameters p. Two selected particles, one passing and another one reaching the spherical surface, are labeled (a) and (b). The angle between the particle trajectory and the connecting wire is not relevant for the analysis, and here, they are shown to be parallel.

Close modal
FIG. 3.

A simple physical model illustration of a vertical cut in Fig. 2, showing the particle a) there. The incoming charge e is given by a small black disk. The distance from the origin of the sphere and the particle at position x(t) is r(t)=p2+x2(t) with p being the impact parameter. A small open circle illustrates the image charge eR/r(t) at a radial position R2/r(t) on the line connecting the original charge e and the center of the sphere. A thick line with an arrow gives the wire carrying the current i(t). For e<0, a positive charge will flow into the sphere in the illustrated wire, giving a negative current i(t) out of the sphere.

FIG. 3.

A simple physical model illustration of a vertical cut in Fig. 2, showing the particle a) there. The incoming charge e is given by a small black disk. The distance from the origin of the sphere and the particle at position x(t) is r(t)=p2+x2(t) with p being the impact parameter. A small open circle illustrates the image charge eR/r(t) at a radial position R2/r(t) on the line connecting the original charge e and the center of the sphere. A thick line with an arrow gives the wire carrying the current i(t). For e<0, a positive charge will flow into the sphere in the illustrated wire, giving a negative current i(t) out of the sphere.

Close modal
FIG. 4.

Illustration of the method of image charges for a spherical conducting surface with radius R. The sphere is at the reference potential, in the sense explained in the text. The central circle is a cut in the sphere as in Fig. 3 and the red point is the imposed external charge, e.g., the charge b) in Fig. 2 taken at some selected time. Solid black curves are equipotentials, and blue curves give the corresponding electric fields. The vertical axis (in arbitrary units) shows the surface charge distribution induced by an external charge e at a distance d from the origin as given by a small red “disk.” The net surface charge on the sphere corresponds to the integral of this charge. The image charge +eR/r at a distance R2/r is illustrated by a smaller blue disk. This is the charge that flows through the connecting wire shown in Figs. 2 and 3.

FIG. 4.

Illustration of the method of image charges for a spherical conducting surface with radius R. The sphere is at the reference potential, in the sense explained in the text. The central circle is a cut in the sphere as in Fig. 3 and the red point is the imposed external charge, e.g., the charge b) in Fig. 2 taken at some selected time. Solid black curves are equipotentials, and blue curves give the corresponding electric fields. The vertical axis (in arbitrary units) shows the surface charge distribution induced by an external charge e at a distance d from the origin as given by a small red “disk.” The net surface charge on the sphere corresponds to the integral of this charge. The image charge +eR/r at a distance R2/r is illustrated by a smaller blue disk. This is the charge that flows through the connecting wire shown in Figs. 2 and 3.

Close modal
When the reference particle moves with a constant impact parameter p, it induces a time variation of q(t) and a corresponding varying current i(t)=dq(t)/dt in the connecting wire so that abi(t)dt=q(a)q(b). Let the particle velocity be a constant w and x=wt. A straight-line orbit is admissible when the conducting surface is at the reference potential, see also the  Appendix. Then,
(1)
and the time-varying current in the connecting wire is
(2)

The time origin is here chosen so that the charge e is at the smallest distance to the origin of the sphere at t=0. The discussion assumed that the change in the charging of the conducting sphere is mediated by currents in a connecting wire illustrated in Fig. 3. With the present current definition and sign convention, we find i(t)>0 for an approaching positive particle, e>0, where t<0 and i(t)<0 for a receding one when t>0.

In the case of a floating sphere, the net charge on the sphere is constant while its potential will be changing as a charged particle approaches. Then, the current i(t) will be compensated by the ε0E/t term (Maxwell's displacement current) in Maxwell's equations. The two problems, floating and grounded/biased spheres, can be treated the same way as far as single-particle motions are concerned. The difference will be manifested when many charges are included: for a floating conductor, the already absorbed charges will affect the motion of those approaching at a later time, while the potential remains fixed for the grounded case.

A charged particle will be accelerated by the image charges it induces in the conducting surface. The assumption of constant particle velocities can thus not be exact. To estimate the error, we compare the initial kinetic particle energy 12mw2 with the energy the particle will attain by interacting with its image charge. Integrating the force F=e2/(4πε0(rR2/r)2), it is found that the work done in bringing the charge e from infinity to the position p is (e2R)/(8πε0(p2R2)), i.e., when the particle arrives at the distance corresponding to the impact parameter p>R in Fig. 3. To compare these two energies, it is assumed that the reference particle is taken from a population with temperature T giving   12mw2T on average. Then,
(3)
Multiplying and dividing by a reference plasma density n, we can formally again introduce the plasma parameter16,22 NpnλD31 and γp/R>1 with λD being the Debye length. For most relevant purposes (except small metallic dust grains, etc.), we have R>λD. The ratio in Eq. (3) will be large, except for cases where the normalized impact parameter γ is close to unity. Consequently, the particle orbit will generally be perturbed only little by its image charge, so w can be taken to be constant as assumed here. As an approximation, we can assume that the motion of a charged particle arriving later is unaffected by its own as well as other image charges. A more detailed discussion is given in the  Appendix.
Assume now that a number N of many independently moving charged particles are present, all with the same velocity and same impact parameter, the only distinction being their random position along the horizontal dashed line in Fig. 3 with constant impact parameter. The collective interactions of the particles can be ignored when their density is small. The interactions with the image charges are ignored as well according to the arguments in Sec. II A and the  Appendix. The time varying current in a long time record becomes
(4)
where each of the i(t)’s is given through Eq. (2). The randomly distributed arrival times tk are taken to be uniformly and independently distributed in a long time interval {T/2;T/2}. The two limits p>R and p<R require different analyses.
For the case of passing particles p>R, the time series Eq. (4) can be analyzed by standard methods17,23,24 (Campbell's theorem) giving the results relevant here as
(5)
(6)
where μ is the density of moving charges in the time record and γp/R. There is no singularity for p0, since p>R. Bypassing charged particles will not give rise to any direct-current but contribute to the root mean square (RMS) noise in the wire that is connected to the sphere. The results Eqs. (5) and (6) refer to ensemble averages and not to individual events or realizations of, for instance, short-time numerical simulations.
For the case where p<R a charged particle such as b) in Fig. 2 will be absorbed by the conducting surface at a time tr=(1/w)R2p2. The current pulse derived from Eq. (2) here becomes
(7)
or
(8)
with Heaviside's unit step function S(t) with ξtw/p and γp/R.

The current i(t) is illustrated in Fig. 5 for a range of p-values. When the charged reference particle reaches the conducting surface, it neutralizes the image charge (or rather the induced surface charge), but it does not give any particular contribution to the current i(t) at that time. The time variation of the current pulse depends on the normal and tangential velocity components of the local surface element.

FIG. 5.

Illustration of the current i(t) given by Eqs. (2) and (7) for a range of impact parameters, p, for a sphere at the reference potential. A contour-plot of i(t) for varying parameters is inserted at the top of the figure. A thin white horizontal line shows p=R for reference. A positive current is flowing out of the probe and the figure shows i(t) for a positive charge arriving from t<0.

FIG. 5.

Illustration of the current i(t) given by Eqs. (2) and (7) for a range of impact parameters, p, for a sphere at the reference potential. A contour-plot of i(t) for varying parameters is inserted at the top of the figure. A thin white horizontal line shows p=R for reference. A positive current is flowing out of the probe and the figure shows i(t) for a positive charge arriving from t<0.

Close modal
Following the discussion of Eq. (4), the average net current value is here found to be
(9)
independent of p and R as long as p<R. The result for the mean square current fluctuation level is more complicated,
(10)
with 0γp/R1. A series expansion of the numerator in Eq. (10) gives 163Rp385p5/R+O(p6), so there is no singularity for p0. It can be shown that 0<F(γ)<1, see Fig. 6. The expression (10) is singular for R0, but pR, so in this limit, both the sphere and the incoming particles will vanish. The relations (9) and (10) demonstrate that current and noise are inseparable. The charged particles give both current and noise.
FIG. 6.

Illustration of the normalized function F(γ) defined for Eq. (10), where γp/R. Limiting values are F(0)=1/3 and F(1)=π/16.

FIG. 6.

Illustration of the normalized function F(γ) defined for Eq. (10), where γp/R. Limiting values are F(0)=1/3 and F(1)=π/16.

Close modal

A complete solution to the problem requires the standard averaging of the distribution of particle velocities and directions for all species as well as all impact parameters p. For relevant conditions, the light species (electrons) will have an RMS velocity exceeding that of the heavy ones and the probe will carry a net current.6 Additional corrections will be induced by inhomogeneities and bulk relative particle motions.

In order to relate the result to measurable quantities, the density μ of signals in the time-record needs to be expressed in terms of plasma parameters. We have
(11)
being the average number of particles crossing a small reference area dA per unit time (see Fig. 1), with n being the plasma density, assumed uniform. With this result we find I(t)=enwdAcos θ, i.e., the result also found by generating a time record by marking charges crossing the reference surface area dA. The integral i(t)dt gives the charge flowing through the connecting wire in Figs. 2 and 3, while i(t) gives details of the time variations. While the integral of i(t) over all |p|R and the distribution of velocities w is complicated, the corresponding result for the ensemble average Eq. (9) is simple.

If the record for I(t) is used for estimating the mean square fluctuations of the current, the result will be different from Eq. (10) because of the missing contribution from Eq. (6). The interpretation of the statistical average of the steady state, or DC, current is correct, even though the interpretation of the individual charge contributions is more complicated than commonly assumed.

For p<R, we have the angle θ=arcsin(p/R) between the sphere's radius vector and the velocity vector of the particle at the time of impact, see also Fig. 1. The variation of i(t) shown in Fig. 5 for varying p can also be seen as representing the change in current for varying θ, i.e., varying direction of the velocity vector at a fixed point of impact on the sphere. When I in Eq. (9) is independent of p and R, it is also independent of θ. The variable γ in F(γ), see Fig. 6, is directly related to θ.

1. Relation to a simple model

Taking p<R, it is readily shown that Eq. (7) taken at the time of impact tr=(1/w)R2p2 gives
while q(tr)=e by Eq. (1). Physically, this means that at the time of impact of the moving charge, the spherical surface can be approximated by a local plane where the image charge is e. This result implies i(t)/edt=1. For all ttr, we have limwi(t)/e0, while limwi(tr)/e. The physical dimension of i(t)/e is time 1. The consistent limit25,26 is then limwi(t)/e=δ(t) in Eq. (4), implying that the usually assumed simplification of “pulse-like” arrivals to the probe surface is acceptable for very fast particles, i.e., better for thermal electrons than for ions at similar temperatures. In the limit w, the problem reduces to that of “shot noise.” The auto-correlation function for this limit is a δ-function and the power spectrum derived from it is “white noise” with a DC component I(τ).

Using the extensions of the Campbell theorem,24,27,28 we can obtain the auto-correlation function for the current I(τ)I(τ+t) for the model (4). Again, the two limits pR and p>R can be identified.

By use of Eq. (2), we have the form
(12)
which accounts for both p<R and pR with the appropriate form for i(t) inserted. Using Eq. (2) for p>R, the expression is found in the form
(13)

Results for the correlations I(τ)I(τ+t) are shown in Figs. 7 and 8 for p>R and p<R, respectively, omitting the (i(τ)dτ)2 constant part for the case pR.

FIG. 7.

Numerical solution for the correlation function i(τ)i(τ+t)dτ for p>R for charged particles passing a sphere at the reference potential. A (i(τ)dτ)2 part is vanishing for this case. For t=0, we recover Eq. (6).

FIG. 7.

Numerical solution for the correlation function i(τ)i(τ+t)dτ for p>R for charged particles passing a sphere at the reference potential. A (i(τ)dτ)2 part is vanishing for this case. For t=0, we recover Eq. (6).

Close modal
FIG. 8.

Correlation functions i(τ)i(τ+t)dτ for the case p<R, i.e., for charged particles reaching the surface of a sphere at the reference potential. Note the change in normalizations compared to Fig. 7. The (i(τ)dτ)2 part is omitted for this case. For t=0, we recover Eq. (10). The correlation has a simple analytical form for p=0.

FIG. 8.

Correlation functions i(τ)i(τ+t)dτ for the case p<R, i.e., for charged particles reaching the surface of a sphere at the reference potential. Note the change in normalizations compared to Fig. 7. The (i(τ)dτ)2 part is omitted for this case. For t=0, we recover Eq. (10). The correlation has a simple analytical form for p=0.

Close modal

The power spectra are obtained by Fourier transforming the auto-correlation functions, with an example shown for p>R in Fig. 9. The spectrum has a broad peak at ω2πw/p for this case. This result is consistent with a characteristic interaction time22 of p/w. Nearby particles (small p) contribute with large frequencies to the net power spectrum and have the largest influence due to the p3 factor. The number of incoming particles for increasing p increases as 2πpdp, implying that the net noise contribution from distant particles reduces as p2. A dominant frequency will be 2πw21/2/R.

FIG. 9.

Power spectrum obtained by Fourier transform of the autocorrelation function in Fig. 7 for p>R. The shape of the spectrum is independent of p, apart from the normalization of the frequency variable, while its intensity decreases as p3. A larger velocity w implies a wider frequency range of the spectrum, a smaller impact parameter p will have the same effect.

FIG. 9.

Power spectrum obtained by Fourier transform of the autocorrelation function in Fig. 7 for p>R. The shape of the spectrum is independent of p, apart from the normalization of the frequency variable, while its intensity decreases as p3. A larger velocity w implies a wider frequency range of the spectrum, a smaller impact parameter p will have the same effect.

Close modal

The system considered here, taken as a whole, is not in thermal equilibrium, so the fluctuation–dissipation theorem24,29,30 does not apply. For the present conditions, the noise induced by image charges will have two distinct contributions from p>R and from pR.

The particle's contribution to the power spectrum of the noise when p<R can be found similarly by Fourier transform of the auto-correlation function in Fig. 8, see Fig. 10 for the result. The shape of the power spectrum depends on the angle of impact with respect to the sphere's local radius vector at the time the particle reaches the surface. The spectra differ significantly from the “white noise” resulting from the classical shot noise17 in both cases p<R and p>R. There appears to be no analytical expression for the power spectra in Fig. 10. An averaging over the distribution of particle velocities and directions for each species as well as all impact parameters has to be performed numerically. The noise spectra for p<R shown in Fig. 10 do not have local maxima.

FIG. 10.

Double logarithmic presentation of the power spectrum obtained by Fourier transform of the auto-correlation function in Fig. 8 for p<R. The particle's contribution to the noise spectrum depends on its angle of impact when it reaches the sphere.

FIG. 10.

Double logarithmic presentation of the power spectrum obtained by Fourier transform of the auto-correlation function in Fig. 8 for p<R. The particle's contribution to the noise spectrum depends on its angle of impact when it reaches the sphere.

Close modal

The analysis assumes that the image charges respond instantaneously to changes in the position of the moving reference charge, i.e., assuming perfectly conducting surfaces. This assumption is expected to work best for slowly moving charges. Some basic features of finite conductivity can be modeled by inserting an impedance in the connecting wire, e.g., a self-inductance for the finite inertia of the conducting electrons in the metal, and a resistor to model the internal collisional friction. The electrical network connected to the probe is essential for its overall performance,4 but this aspect is not addressed here.

The analysis presented in the foregoing Sec. II assumed straight-line orbits of the charged particles, implying that the spherical surface is at the plasma potential, denoted “ground” or reference potential before. A DC bias of the probe different from ground will give rise to electric fields that will perturb the particle orbits1,6 to give r(t)=tu(τ)dτ in q(t)=eR/r(t)·r(t) and replacing Eq. (2) by
(14)
where r(t) is the vector connecting the origin of the conducting sphere and the reference particle. The expression (2) is a special case with r2(t)=p2+w2t2. The full analysis for biased probes need to be carried out by considering individual orbits.

A bias on the sphere at a potential Φ different from the reference ground can be modeled by introducing a charge Q=4πε0RΦ at the center. This charge will create an electric field that perturbs the orbits of an approaching charge. At large distances rR, the resulting force will in general dominate the contribution from the image charge. The straight-line orbits in Figs. 2 and 3 will then be changed31 into parabolic or hyperbolic orbits in Eq. (14). For simple symmetries and geometries, also this more general problem may be solvable, but the general case requires a numerical analysis.

Sections II and III addressed spherical surfaces at a fixed potential with a fluctuating current. An alternative, and also relevant problem, deals with a floating conductor where the potential fluctuates and the net average current is vanishing, I(t)=0. For reasons outlined in the following, this is a more complicated problem. The fluctuating potential gives here rise to a fluctuating electric field, which gives ε0E/t0 in Ampère's law to compensate the current to the surface caused by the randomly arriving charged particles.

The problem of a single particle arriving at a floating charge neutral sphere can be analyzed along the same lines as for the grounded or biased spherical conductor. For a grounded probe, the image charge will make the surface an equipotential. For the corresponding floating problem, an additional charge eR/r similar in magnitude but with opposite sign has to be introduced at the center to make the sphere charge neutral. This central charge taken alone will make the spherical surface equipotential. The surface potential will vary with time in case the reference particle is moving. An illustrative solution of the problem is given in Fig. 11. The two image charges inside the spherical surface form a dipole displaced from the center: a somewhat counter-intuitive result. A passing particle with charge e, velocity w and impact parameter p gives rise to a time varying potential ϕ(t)=e/(4πε0p2+w2t2) at the surface of the sphere, independent its radius. The magnitude of the dipole moment Di is the product of the image charge and the separation distance between the inner charges in Fig. 11, i.e., Di(t)=|e|R3/(p2+w2t2). The direction of the dipole moment vector varies from π to π during the passage of the charged particle.

FIG. 11.

Illustration of the electric field and potential variation for an unbiased or floating spherical conductor. Compare with Fig. 4. The surface is shown here with a thick dashed line. A large red disk represents the external charge. The two image charges ±eR/d of opposite polarity are positioned at the origin and at a distance R2/d inside the sphere. This configuration ensures that the floating sphere is charge neutral. The incoming charge together with the image charges make the sphere equipotential ϕ(t)=e/(4πε0p2+w2t2) at all times.

FIG. 11.

Illustration of the electric field and potential variation for an unbiased or floating spherical conductor. Compare with Fig. 4. The surface is shown here with a thick dashed line. A large red disk represents the external charge. The two image charges ±eR/d of opposite polarity are positioned at the origin and at a distance R2/d inside the sphere. This configuration ensures that the floating sphere is charge neutral. The incoming charge together with the image charges make the sphere equipotential ϕ(t)=e/(4πε0p2+w2t2) at all times.

Close modal
For a biased surface, it is feasible to analyze many particle arrivals individually. It is not so for a floating surface embedded in a plasma. In addition to the fluctuating electric field component induced by the reference particle, the sum of the previously arrived particles will give rise to an average potential for floating surfaces. This average potential will give an average electric field that influences the trajectories of particles arriving later. At the saturated stage, the sphere will assume the usual average floating potential6,31,32 with an additional fluctuating component. When TeTi, a commonly accepted approximation7,33 for the floating potential Φf taken with respect to the reference ground or plasma potential for a thermal plasma is
(15)
The expression takes into account the Bohm condition,22,33,34 i.e., that ions arrive at the sheath edge with the sound velocity. At the saturated plasma, potential most electrons, except the most energetic ones, are reflected, while all incoming ions are absorbed at the surface. The average current to a floating sphere will vanish, but there is no detailed balance. In particular, the surface potential will be varying also due to passing charges. Compared to the average field, the fluctuating electric fields will usually only give small modifications of the incoming particle trajectories, see also the discussion in the  Appendix.

When more particles are introduced, it will be found that the arrival of the next one will be influenced by the potential induced by the previous ones, etc., as discussed before. Campbell's theorem is no longer directly applicable for the present problem since the particle arrivals can no longer be considered independent. The process is no longer time stationary. Asymptotically, when the sphere is saturated at the floating potential, it will again be possible to assume stationary processes. The analysis of this many particle problem is likely to be carried out best by a combination of analytical and numerical methods.

The presence of dipolar charge distributions has been discussed and illustrated by numerical simulations also in relation to conducting dust particles moving at high speeds relative to a fully ionized plasma.35 In such cases, a permanent dipole will be due to an asymmetrical distribution of the charges absorbed on the surface of the conducting dust grain or a similar finite size object.

The collisional case differs notably from the collisionless case discussed in Sec. II. The current I(t) will remain a stationary process, but the participating plasma particles can no longer be distinguished in two types as in Fig. 2. Passing particles like a) in Fig. 2 will be scattered and can reach the conducting surface, while particles of type b) can be scattered out of orbit to be lost eventually at r, never to reach the conducting sphere. Between two collisions, the signal can be presented by a segment of Eq. (2) with a suitably chosen impact parameter. To illustrate the collisional case, we consider an unmagnetized particle undergoing only one collision at some time T. In the time interval t<T, we have q(t)=eRS(Tt)/p2+w2t2 introducing again Heaviside's unit step function, the signal becomes
(16)
At the time of collision, we let this particle disappear to be replaced at the same time and position with a new scattered particle with an impact parameter p1 and velocity w1. This particle is assumed to either (i) continue to infinity or (ii) be lost to the spherical surface. The case with intermediate collisions brings nothing new to the discussion.
Taking case (i) first, we have for t>T1,
(17)
Here, a new time-axis t1 was introduced, so that at the moving particle has its minimum distance p1 to the sphere at time t1=0. The time of collision is t1=T1 with this new time-axis. For this case, the times T and T1 can assume either sign by the present definition of the time-axes. The vector w1 need not be in the plane spanned by w and the center of the conducting sphere. The current will be discontinuous at the time of collision, the charge continuous. Irrespective of the definition of the time axes, we have Ti(t)dt=0 and T1i1(t1)dt1=0, implying that the moving charge will not contribute to the average current for case (i). The correlation function of the current will contain a δ-function due to the δ-functions in Eqs. (16) and (17), giving rise to a white noise contribution to the power spectra. Physically, this white noise is due to the assumed vanishing time-duration of the collision.
For the case (ii), the expression (17) is replaced by
(18)
For case (ii), we have T1<0, and otherwise, the particle has already passed the sphere at this time. The charge e does not disappear when arriving at the spherical surface, it merely cancels the image charge. With Eq. (18), we find the integral T1(1/w1)R2p12i1(t1)dt1=e. Consequently, the result I(t)=μe holds for the collisional as well as the collisionless case (ii).

The general problem with many collisions is studied best by numerical simulations, with results depending critically on the collisional model applied. The average currents remain the same also in collisional plasmas, but the power spectrum of the noise can be significantly different from the collisionless case.

So far, the analysis did not impose any restrictions on the impact parameter p. For relevant plasma conditions, it will be argued that charges with p>λD will be shielded and not generate any image charges in the conducting surface. Considering one spherical probe with radius R as a reference case two limits can be distinguished, sheath limitation where the sheath is narrow compared to R and orbit limitation36 where the sheath is wider, having in mind6 that the sheath can be somewhat wider than λD. A commonly accepted sheath-width is Dsw15λD, depending weakly on applied voltages.6,34 In the former case with R much larger than the sheath-width, almost all incoming charges will reach the conducting surface and the contribution (9) is dominant. For the latter orbit limited case,31 the fluctuating currents Eq. (2) induced by passing particles can be noticeable. Using a simple, and often applied model, we can assume that charges more than a sheath-width from the conducting surface are completely shielded and has no effect the probe. Charges that are closer will have an effect almost as in vacuum. The interaction between the charge and its image charge is restricted to the time spent inside the sheath. The effective time of interaction is estimated by ΔtDsw/w. For large probes, the current signal will only be a fraction of the time variation shown in Fig. 5.

The effects of the sheath can be modeled by formally introducing a charge moving with w>0 at the sheath edge at time T(1/w)(R+Dsw)2p2<0 with 0<p<R+Dsw, so that q(t)=e(R+Dsw)S(tT)/p2+w2t2. The remaining part of the discussion becomes similar to the collisional problem, and again, we find the result I(t)=μe, while the correlation function (and thereby the power spectrum of the noise) becomes complicated, amenable only to a numerical solution.37 

A biased sphere can be modeled as discussed in Sec. III. A probe with a swept bias can be analyzed in the same way as one at a fixed potential provided the sweep is slow when measured on a timescale of Dsw/w. Then, the sheath changes only little during a particle transit, and the problem can be analyzed as a series of quasi-stationary conditions. A fast sweep of the bias is likely to require a numerical approach.

The expression (14) can be used also for a magnetized reference particle. Figure 12 serves as an illustration for a trajectory that deviates from a straight line. If the gyrocenter coincides with the center of the conducting spherical surface, we have r(t)·u(t)=0 and no steady current is induced when r>R. For the more general case, we introduce r(t)=R0(t)+r(t) with R0(t) being a vector from the center of the conducting sphere to the gyrocenter, r(t) is a vector from the gyrocenter to the particle, and u=v+wx̂, where vB while also vr. Then, r(t)·u(t)=R0(t)·u(t)=pωceζsin(ωcet)+w2t, and r2(t)=R02(t)+ ζ2+2R0(t)·r(t)=R02(t)+ζ2+2pζcos(ωcet), with the Larmor radius being ζ=v/ωce, where ωce is the gyro-frequency and v the particle's gyro-velocity, while w is the B-parallel velocity component so that |R0(t)|=p2+w2t2, noting also that projections of r and u are π/2 out of phase. An initial gyro-phase is ignored. The expressions summarized here are to be inserted into Eq. (14) to give the full expression for i(t). The magnetized problem is complicated by having two independent time scales 1/ωce and R/w, as well as two length scales p and ζ, while p/R is a dimensionless parameter. We have
(19)
FIG. 12.

Magnetized version of Fig. 2 where the Larmor radius ζ+R<p. The spiral orbit of the magnetized charge is shown by a dashed line.

FIG. 12.

Magnetized version of Fig. 2 where the Larmor radius ζ+R<p. The spiral orbit of the magnetized charge is shown by a dashed line.

Close modal

It is an advantage to use normalized variables and parameters as p/R, ζ/R, ωceR/w, and wt/R. The expression (19) reproduces Eq. (2) for ζ=0, where the gyrocenter coincides with the particle position and the motion is along a straight line given by B. The special case for p=0 and ζ>R is independent of ωc.

An illustrative example for i(t) given by Eq. (19) is shown in Fig. 13. It can be shown that I(t)=0 also for this case, while I2(t) does not have a simple analytical expression. The signal contains the cyclotron frequency, but it is rich also in its harmonics. The figure assumes, as mentioned, the initial gyro-phase to vanish. This phase is however important for large velocities w when R+ζ>p since in this case, some particles can avoid absorption, see Fig. 14. Here, the contribution to the net current will be vanishing, but the contribution to the RMS-fluctuation level remains. For small w, all particles will be absorbed in this case, but the position on the surface where this happens will depend on the initial phase. Even if the moving charge misses the sphere, it can give a significant contribution to the noise by coming close to its surface.

FIG. 13.

Normalized induced current i(t) for a magnetized charged particle with ζ/R=1/2 and ωceR/w=15/2. The particles are here passing a spherical surface at the reference potential. The figure includes only p>R+ζ. Charges with smaller impact parameter can be absorbed on the conductor at a time depending also on the phase of the gyro-motion.

FIG. 13.

Normalized induced current i(t) for a magnetized charged particle with ζ/R=1/2 and ωceR/w=15/2. The particles are here passing a spherical surface at the reference potential. The figure includes only p>R+ζ. Charges with smaller impact parameter can be absorbed on the conductor at a time depending also on the phase of the gyro-motion.

Close modal
FIG. 14.

Magnetized version of Fig. 2, here with ζ+R>p. Compare with Fig. 12. The particle velocity w and the initial phase are chosen so that the particle “misses” the sphere. Different choices of these parameters will result in an encounter for these values of ζ and p.

FIG. 14.

Magnetized version of Fig. 2, here with ζ+R>p. Compare with Fig. 12. The particle velocity w and the initial phase are chosen so that the particle “misses” the sphere. Different choices of these parameters will result in an encounter for these values of ζ and p.

Close modal

As for the unmagnetized case, we can also here construct the correlation functions of signals like Eq. (19) to the fluctuation spectra detected by the biased spherical probe. This problem has many parameters so an illustration is given only for a selected case where p/R=1.75, ζ/R=0.5, and ωcR/w=5.0, see Fig. 15. The conspicuous feature is the cyclotron harmonics in the power spectrum. Harmonic generation is often associated with physically nonlinear effects, but in the present case, it is a consequence of simple geometrical effects.

FIG. 15.

Sample illustration of the normalized auto-correlation function and the power-spectrum (double logarithmic scale) for the magnetized case of a biased spherical conducting surface, see Fig. 13 for details.

FIG. 15.

Sample illustration of the normalized auto-correlation function and the power-spectrum (double logarithmic scale) for the magnetized case of a biased spherical conducting surface, see Fig. 13 for details.

Close modal

The rare event where a charged particle is trapped in a circular orbit around a biased spherical probe will not give rise to an induced current, while an elliptical orbit will.

The present analysis gives details of the current contributions of charges arriving at the surface of spherical conductors. When compared to simple models where arriving charges are assumed to contribute by δ-pulses, it is found here that the signals become more complicated, depending on the angle of incidence. Particular attention is given to the noise associated with the discrete particle arrivals and it is demonstrated that also passing particles contribute here. For the simple model summarized in Sec. I, they would not contribute at all.

For time stationary homogeneous conditions for a probe at the reference potential, we can estimate the relative importance of the contribution from p<R to the fluctuations I2(t) by using a combination of Eqs. (9)–(11) to consider the ratio
(20)
apart from a numerical factor of order unity. Large collecting areas give a reduced relative noise level.

For scales comparable to or larger than the Debye length λD, we have RdAλD3 in Eq. (20). In that case, I2(t)/I(t)21/Np1 in terms of the plasma parameter Np introduced before. An RMS value for the relative fluctuations scales as Np1/2. For a typical Q-machine plasma8 with Te=2.200 K and n=1015 m−3, we have Np300 for comparison. Due to the density dependence of λD, the plasma parameter is large for hot and dilute plasmas. The arguments outlined here can be repeated using Eq. (10) instead of Eq. (6) giving the same conclusions. In the “Vlasov limit” of large plasma parameters, we can expect contributions Eqs. (6) and (10) to be small but important by setting a minimum noise level for the measurements. For conditions in space plasmas as well as in related numerical simulations with metallic dust grains, it can be relevant to include conducting surfaces on scales smaller than the Debye length, and in that case, the ratio I2(t)/I(t)2 need not be small. All charges contribute to the noise. When a Langmuir probe is used for identifying a subset of particles, such as a small density plasma beam component, it can happen that the signal is masked by the noise caused by all the other particles.

We estimate the parameters for conditions where passing charges give dominant contributions compared to those absorbed. The sheath width will be taken to be simply λD, ignoring numerical coefficients. For a grounded metallic object, the integrated noise contribution from the two regions p<R and R+λD>p>R can be compared by use of Eqs. (6), (10), and (11) giving a mean square noise ratio (1+R/λD) ignoring some numerical factors and taking F(γ)1/4, see Fig. 6. The noise induced by particles reaching the surface will dominate when RλD, while the passing particle noise will become comparable when RλD. With the possible exception of some space applications, plasma probes will usually have R>λD.

The discussions in the preceding sections apply to one conducting surface. Taking as an example a double probe or a combination of two or more surfaces close together,15,38 it will be found that a moving charge contributes to the signal in several of them, although it is ultimately absorbed by only one. A Langmuir probe with a guard-ring5,39 is another example. Two nearby probes can be used to estimate an electric field component.9,15,40,41 Prior to absorption on one surface, a moving charged particle will induce currents also in the other probes with an intensity depending on the relevant impact parameter and probe separations. There seems to be no analytical solution to this formulation of the problem, in particular also because image charges induced by a moving charge in one surface induce image charges in the other surfaces, etc.

The foregoing discussion can be extended to other probe forms. In some cases, the analysis can be simplified by use of conformal mappings. The case of a large plane probe needs, however, special attention.42 

Addition of a nearby insulating surface is in principle straightforward. In this case, it is physically realistic that charges are accumulating there and these will induce increasing image charges on the conductor. A steady increase in the accumulated charge jqj=Q(t)t will give rise to a current dQ(t)/dt const. to the conducting sphere, although no charges actually reach the metallic surface.

Contamination of a probe surface43 need not give problems for the analysis as long as it contributes with a uniform conducting layer. The only change will be in the work function at the surface. A layer with varying thickness and composition over the surface is problematic. An imposed bias means that the Fermi level of the conductor is constant and a varying composition implies that the work function is then varying over the surface, so it is no longer equipotential. The present results will have to be modified, but the basic ideas of currents induced by image charges remain applicable. The distinction between absorbed and passing charge contributions can also be maintained. Coatings by insulating materials would have no effect for the first charged particle arriving. It will bind an image charge that gives rise to a corresponding current through the connecting wire in Figs. 2 and 3. Charges arriving later will pile up until an equilibrium is reached, making the surface potential different from the one imposed by the external connection. The analysis will be reminiscent of the one for a floating conducting surface.

The effects of passing particles can be tested experimentally by placing the spherical probe in a glass tube that cancels the direct current. Results from different tube diameters can be compared to test the effects of varying impact parameters. For instance, a density modulated narrow collimated beam of charged particles passing by the probe will induce a time varying current in the wire connected to the probe.42 The signal will be indistinguishable for a similar case in which the particles reach the probe surface. The difficulty in the experiment comes from the time-varying accumulation of charges on the outer glass surface. An externally imposed magnetic field can reduce the charging and allow also a qualitative comparison with Fig. 15, keeping in mind30 that harmonic generation can also be due to plasma gradients B0 that give perturbations of the cyclotron orbits.

Parts of Figs. 4 and 11 are based on a program made public by Masatsugu Sei Suzuki, Department of Physics, SUNY at Binghamton.

The author has no conflicts to disclose.

Hans L. Pécseli: Conceptualization (equal); Formal analysis (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

The method of images rests on a uniqueness theorem valid for electrostatics. In principle, it breaks down when the charges are in motion. It is argued in the following that the method can be applied as a good approximation for realistic plasma conditions.

A reference charge moving with constant velocity can be approximated by letting it move step-by-step along a prescribed trajectory from one stationary position to the next. As it approaches the grounded conducting sphere, an increment of image charge has to flow into the sphere for each step, thus representing the current through the conducting wire, and the method of images applies at each step. For modeling a freely moving charge, this procedure ignores the interaction through electrical forces between the reference and the image charges. This force gives rise to an acceleration of the incoming charge and a resulting radiation of electromagnetic waves.44 In the following, it is argued that this process is negligible for most relevant problems. Referring to Fig. 3, the electric force equation is found in the form
(A1)
to be written in normalized units as
(A2)
using ρr/R, τtωp, and K4πNp(R/λD)3. By these normalizations, it is assumed that the reference particle with mass m is taken from a population with density n and temperature T giving the appropriate plasma frequency and Debye length. For relevant cases, we have K1. By inspection of Eq. (A2), it is found that the right-hand side is of the order unity. Consequently, a large K-value gives a small acceleration d2ρ(τ)/dτ2, resulting in a nearly constant velocity given by the initial conditions.

Numerical solutions of Eq. (A2) are shown in Fig. 16 for selected parameters K. For small K, the particle can be attracted by its image charge to reach the conducting surface. However, for the given impact parameter and reference velocity, even moderate parameters as K>15 give only small deflections of the particle. For K100, the accelerations are negligible and the assumption of a constant velocity w used in the main analysis is justified. When the charged particle moves without acceleration, it will not give rise to any electromagnetic radiation. When K1, the use of the method of images will be justified also for the present dynamic problem.

FIG. 16.

Particle trajectories with ρ(t)=x2(t)+y2(t) in (A2) for selected parameters, here K=50 with full line, K=25 dash-dotted line, and K=15 dashed line. The impact parameters are p=1.5 (red curves) and 1.25 (blue curves). In all cases, the initial velocity is w=1 in normalized units. A semi-circle shown with a thick black line gives a cross section of the conducting sphere. The choice R=λD makes w=1 corresponding to the thermal velocity in physical units.

FIG. 16.

Particle trajectories with ρ(t)=x2(t)+y2(t) in (A2) for selected parameters, here K=50 with full line, K=25 dash-dotted line, and K=15 dashed line. The impact parameters are p=1.5 (red curves) and 1.25 (blue curves). In all cases, the initial velocity is w=1 in normalized units. A semi-circle shown with a thick black line gives a cross section of the conducting sphere. The choice R=λD makes w=1 corresponding to the thermal velocity in physical units.

Close modal

A floating conductor that has reached a saturated floating potential given by Eq. (15) can be modeled by introducing a charge Qf=4πε0RΦf at the center of the sphere. When the external moving charge comes very close to the conducting surface, the electric forces on it will be dominated by the image charge. At larger distances, however, the force from the dipole in Fig. 11 decreases as r3 compared to the r2 force from Qf. In this case, the image charge force can be ignored under all circumstances, retaining only the force from Qf.

In the past, it has been speculated that charged particles could be attracted by metallic dust particles in space and by their acceleration contribute to low energy electromagnetic background radiation. The idea may be worth pursuing, but it needs an approach extending the one summarized here. For instance, the radiation losses44 of the accelerated charges will need attention. Phenomenological models for this can be found in the literature.24,44

1.
J. G.
Laframboise
, “
Theory of spherical and cylindrical Langmuir probes in a collisionless, Maxwellian plasma at rest
,” Report No. 100 (
University of Toronto, Institute for Aerospace Studies
,
1966
).
2.
F. F.
Chen
, “
Saturation ion currents to Langmuir probes
,”
J. Appl. Phys.
36
,
675
678
(
1965
).
3.
A.
Skøelv
,
R. J.
Armstrong
, and
J.
Trulsen
, “
Ion-beam diagnostics by means of plane Langmuir probes
,”
Phys. Fluids
27
,
2744
2751
(
1984
).
4.
V. I.
Demidov
,
S. V.
Ratynskaia
, and
K.
Rypdal
, “
Electric probes for plasmas: The link between theory and instrument
,”
Rev. Sci. Instrum.
73
,
3409
3439
(
2002
).
5.
F. F.
Chen
, “
Langmuir probe diagnostics
,” in
IEEE-ICOPS Meeting, Jeju, Korea
(
Citeseer
,
2003
), Vol.
2
.
6.
M. A.
Lieberman
and
A. J.
Lichtenberg
,
Principles of Plasma Discharges and Materials Processing
, 2nd ed. (
John Wiley & Sons, Inc
.,
New York
,
2005
).
7.
R. L.
Merlino
, “
Understanding Langmuir probe current-voltage characteristics
,”
Am. J. Phys.
75
,
1078
1085
(
2007
).
8.
R. W.
Motley
,
Q Machines
(
Academic Press
,
New York
,
1975
).
9.
N.
Yan
,
A. H.
Nielsen
,
G. S.
Xu
,
V.
Naulin
,
J. J.
Rasmussen
,
J.
Madsen
,
H. Q.
Wang
,
S. C.
Liu
,
W.
Zhang
,
L.
Wang
, and
B. N.
Wan
, “
Statistical characterization of turbulence in the boundary plasma of EAST
,”
Plasma Phys. Controlled Fusion
55
,
115007
(
2013
).
10.
F. F.
Chen
, “
Numerical computations for ion probe characteristics in a collisionless plasma
,”
J. Nucl. Energy. Part C, Plasma Phys.
7
,
47
67
(
1965
).
11.
R. J.
Armstrong
,
K. B.
Liland
, and
J.
Trulsen
, “
Plasma simulation of a Langmuir probe with normal magnetic-field
,”
Phys. D
60
,
160
164
(
1992
).
12.
D.
Darian
,
S.
Marholm
,
M.
Mortensen
, and
W. J.
Miloch
, “
Theory and simulations of spherical and cylindrical Langmuir probes in non-Maxwellian plasmas
,”
Plasma Phys. Controlled Fusion
61
,
085025
(
2019
).
13.
M.
Komm
,
J.
Adamek
,
J.
Cavalier
,
J.
Brotankova
,
O.
Grover
,
J.
Hecko
,
J.
Horacek
,
J.
Matejicek
,
M.
Peterka
,
A.
Podolnik
,
J.
Seidl
,
M.
Hron
, and
R.
Panek
, “
On the applicability of three and four parameter fits for analysis of swept embedded Langmuir probes in magnetised plasma
,”
Nucl. Fusion
62
,
096021
(
2022
).
14.
K. M.
Kjølerbakken
,
W. J.
Miloch
, and
K.
Røed
, “
Sheath formation time for spherical Langmuir probes
,”
J. Plasma Phys.
89
,
905890102
(
2023
).
15.
C.-S.
Jao
,
W. J.
Miloch
, and
Y.
Miyake
, “
Collected current by a double Langmuir probe setup with plasma flow
,”
IEEE Trans. Plasma Sci.
52
,
5222
5233
(
2024
).
16.
H. L.
Pécseli
,
Waves and Oscillations in Plasmas
, 2nd ed. (
Taylor & Francis
,
London
,
2020
).
17.
W. B.
Davenport
and
W. L.
Root
,
An Introduction to the Theory of Random Signals and Noise
(
McGraw-Hill
,
New York
,
1958
).
18.
J. A.
Stratton
,
Electromagnetic Theory
(
McGraw-Hill Book Company
,
1941
).
19.
D. J.
Griffiths
,
Introduction to Electrodynamics
, 3rd ed. (
Prentice Hall
,
Englewood Cliffs, NJ
,
1999
).
20.
W. T.
Scott
,
The Physics of Electricity and Magnetism
(
John Wiley & Sons
,
New York
,
1959
).
21.
J.
Kunz
and
P. L.
Bayley
, “
Some applications of the method of images–I
,”
Phys. Rev.
17
,
147
156
(
1921
).
22.
F. F.
Chen
,
Introduction to Plasma Physics and Controlled Fusion
, 3rd ed. (
Springer
,
Heidelberg
,
2016
).
23.
N.
Campbell
, “
The study of discontinuous phenomena
,”
Proc. Cambridge Philos. Soc.
15
,
117
136
(
1909
); ibid. 15, 310–328 (1909).
24.
H. L.
Pécseli
,
Fluctuations in Physical Systems
(
Cambridge University Press
,
Cambridge, UK
,
2000
).
25.
M. J.
Lighthill
,
An Introduction to Fourier Analysis and Generalized Functions
(
Cambridge University Press
,
London
,
1964
).
26.
E.
Berz
,
Verallgemeinerte funktionen und operatoren
(
Hochschultaschenbücher, Bibliographisches Institut
,
Mannheim
,
1967
), Vol.
122/122a
.
27.
S. O.
Rice
, “
Mathematical analysis of random noise, I
,”
Bell Syst. Tech. J.
23
,
282
332
(
1944
).
28.
S. O.
Rice
, “
Mathematical analysis of random noise, II
,”
Bell Syst. Tech. J.
24
,
46
156
(
1945
).
29.
D. K. C.
MacDonald
,
Noise and Fluctuations: An Introduction
(
John Wiley & Sons
,
New York
,
1962
).
30.
G.
Bekefi
,
Radiation Processes in Plasmas
(
John Wiley and Sons
,
New York
,
1966
).
31.
R. V.
Kennedy
and
J. E.
Allen
, “
The floating potential of spherical probes and dust grains. II: Orbital motion theory
,”
J. Plasma Phys.
69
,
485
506
(
2003
).
32.
R. V.
Kennedy
and
J. E.
Allen
, “
The floating potential of spherical probes and dust grains. Part 1. Radial motion theory
,”
J. Plasma Phys.
67
,
243
250
(
2002
).
33.
A.
Piel
,
Plasma Physics: An Introduction to Laboratory, Space, and Fusion Plasmas
, 2nd ed. (
Springer
,
Heidelberg
,
2017
).
34.
M.
Kono
and
H. L.
Pécseli
, “
Dynamic evolutions of Bohm sheaths and pre-sheaths
,”
Phys. Plasmas
31
,
023508
(
2024
).
35.
W. J.
Miloch
,
H. L.
Pécseli
, and
J.
Trulsen
, “
Numerical simulations of the charging of dust particles by contact with hot plasmas
,”
Nonlinear Process. Geophys.
14
,
575
586
(
2007
).
36.
J. G.
Laframboise
and
L. W.
Parker
, “
Probe design for orbit‐limited current collection
,”
Phys. Fluids
16
,
629
636
(
1973
).
37.
I.
Holod
and
Z.
Lin
, “
Statistical analysis of fluctuations and noise-driven transport in particle-in-cell simulations of plasma turbulence
,”
Phys. Plasmas
14
,
032306
(
2007
).
38.
M. Y.
Naz
,
A.
Ghaffar
,
N. U.
Rehman
,
S.
Naseer
, and
M.
Zakaullah
, “
Double and triple Langmuir probes measurements in inductively coupled nitrogen plasma
,”
Prog. Electromagn. Res.
114
,
113
128
(
2011
).
39.
D. L.
Brown
,
M. L. R.
Walker
,
J.
Szabo
,
W.
Huang
, and
J. E.
Foster
, “
Recommended practice for use of Faraday probes in electric propulsion testing
,”
J. Propul. Power
33
,
582
613
(
2017
).
40.
S. J.
Zweben
, “
Search for coherent structure within tokamak plasma turbulence
,”
Phys. Fluids
28
,
974
982
(
1985
).
41.
T.
Huld
,
S.
Iizuka
,
H. L.
Pécseli
, and
J. J.
Rasmussen
, “
Experimental investigation of flute-type electrostatic turbulence
,”
Plasma Phys. Controlled Fusion
30
,
1297
1318
(
1988
).
42.
H. L.
Pécseli
,
Low Frequency Waves and Turbulence in Magnetized Laboratory Plasmas and in the Ionosphere
(
IOP Publishing
,
UK
,
2016
).
43.
C. T.
Steigies
and
A.
Barjatya
, “
Contamination effects on fixed-bias Langmuir probes
,”
Rev. Sci. Instrum.
83
,
113502
(
2012
).
44.
P. C.
Clemmow
and
J. P.
Dougherty
,
Electrodynamics of Particles and Plasmas
, Addison-Wesley Series in Advanced Physics
(Addison-Wesley
,
Reading, MA
,
1969
).