Open Access
22 September 2018 Momentum predictability and heat accumulation in laser-based space debris removal
Stefan Scharring, Lukas Eisert, Raoul-Amadeus Lorbeer, Hans-Albert Eckel
Author Affiliations +
Abstract

Small space debris objects of even a few centimeters can cause severe damage to satellites. Powerful lasers are often proposed for pushing small debris by laser-ablative recoil toward an orbit where atmospheric burn-up yields their remediation. We analyze whether laser-ablative momentum generation is safe and reliable concerning predictability of momentum and accumulation of heat at the target. With hydrodynamic simulations on laser ablation of aluminum as the prevalent debris material, we study laser parameter dependencies of thermomechanical coupling. The results serve as configuration for raytracing-based Monte Carlo simulations on imparted momentum and heat of randomly shaped fragments within a Gaussian laser spot. Orbit modification and heating are analyzed exemplarily under repetitive laser irradiation. Short wavelengths are advantageous, yielding momentum coupling up to ∼40  mNs  /  kJ, and thermal coupling can be minimized to 7% of the pulse energy using short-laser pulses. Random target orientation yields a momentum uncertainty of 86% and the thrust angle exhibits 40% scatter around 45 deg. Moreover, laser pointing errors at least redouble the uncertainty in momentum prediction. Due to heat accumulation of a few Kelvin per pulse, their number is restricted to allow for intermediate cooldown. Momentum scatter requires a sound collision analysis for conceivable trajectory modifications.

1.

Introduction

1.1.

Motivation

When one discovers a shooting star at night, this usually means an opportunity to make a secret wish. For astronauts, this moment means that one wish has already come true: whatever could have harmed or even terminated their mission, burns up now in the atmosphere instead. Man-made space junk meanwhile poses a serious threat to any space mission—being it manned space flight or one of the numerous satellites whose usage has become part of our daily life.

It was already impressively illustrated in April 2010 how small particles may have a significant impact on modern transportation: volcanic ash from Eyjafjallajökull on Iceland posed a massive danger to aircraft engines and thus disabled any aircraft movement over Europe for several days. Likewise, following Kessler and Cour-Palais,1 the exponentially increasing number of space debris might cause such a pollution, referred to as Kessler-syndrome, that certain Earth orbits could turn useless for space flight within only a few decades.

Research and development of laser-based detection and tracking of small-sized debris objects, i.e., in the size range down to 1 cm, is a necessary prerequisite of laser-based removal, but also has a great benefit in itself. The catalog of trajectories of debris objects contains only a few objects <10  cm, whereas most of them are undiscovered, cf. Table 1. Nevertheless, if >1  cm, debris objects typically pose a lethal threat for satellites and manned spacecraft, where no obstacle avoidance maneuver can be provided. Comparing the respective numbers of threatening objects with the number of cataloged and publicly available data, at present Space Situational Awareness resembles a weather forecast restricted on hurricanes neglecting tornadoes, which are orders of magnitude more frequent. In this regard, lasers can not only be applied for space debris monitoring, but also their scalability with respect to average power enables their usage as a tool for debris removal as well.

Table 1

Space debris statistics, taken from Ref. 2.

Debris sizeNumber of objectsProperties
1  mm200,000,000Possible damage to spacecraft
1  cm700,000Satellite wall penetration
10  cm29,000Spacecraft destruction
5  cm16,300Catalogued orbital data

1.2.

Space Debris Targets

Two elaborate models exist for the simulation of the space debris population, namely the Ordem model from the National Aeronautics and Space Administration (NASA) and the Master model from the European Space Agency (ESA).3 Although in the Ordem model debris objects are categorized by their density, the great variety of debris objects considered in the Master model is grouped by its source of formation. This comprises fragments from explosions and collisions, sodium-potassium (NaK) droplets, slag from solid rocket motor (SRM) firing, SRM dust, multilayered insulation, paint flakes, ejecta, and clusters. According to the Master model,4 explosions and collisions are the main source of space debris in the size range from 1 to 10 cm for the sun-synchronous orbits (SSO), where the spatial number density of debris objects >1  cm peaks at almost 106km3. Hence, we focus our considerations on aluminum fragments, which constitute the most prevalent material of this object class.5 Nevertheless, simulations for this size range might also address both fragments of other materials, e.g., steel, copper, phenolic/plastic, and fiberglass, as well as multilayered insulation, NaK droplets, and SRM slag.

1.3.

Laser-Based Remediation Strategies

The orbital velocity is a crucial parameter that determines the orbital trajectory of a space asset. Correspondingly, the applied velocity increment Δv is a central figure of merit in any kind of laser-based debris removal concept. Δv is closely related to another figure of merit in this field, namely the momentum coupling coefficient cm of laser propulsion, which is the analog to debris removal, however, with cooperative targets that are designed for the purpose of being propelled by high-power laser beams.

cm comes with a variety of definitions of which cm=Δp/EL, with Δp as the imparted momentum change and EL as the applied laser pulse energy is the most intuitive one. For our purposes, it is more practical to normalize the defining equation by the target area AL that is cross sectional with the laser-beam yielding Δv=cm·Φ, where Φ is the laser fluence. The strong dependencies of cm on Φ and other parameters like pulse length τ, wavelength λ, incidence angle ϑ, and the target material give rise to a great field of studies of which some issues are treated in the following.

In laser-based debris removal, the underlying physical process is of ablative nature, i.e., based on the recoil of a small fraction of the target surface that is ablated by the incoming high-intensity laser radiation. Due to the superior beam quality of laser radiation from appropriate devices, a focal spot with <1-m in diameter can be achieved even over a distance of several hundreds of kilometers. Therefore, laser-based concepts of space debris removal have been formulated and investigated for both Earth-based as well as satellite-borne laser operations for several decades now.611

In many of those studies, the specific target geometry has been neglected though it has a significant impact on the amount and direction of the laser-induced velocity increment. Therefore, we analyzed laser-induced momentum coupling on irregularly shaped targets in our previous paper.12 Meanwhile, also an experimental validation of those numerical calculations has been carried out with realistic, cm-sized target geometries.13 The main outcome of our analyses was that a notable fraction of the imparted momentum can be found in the plane perpendicular to the laser-beam propagation axis, cf. Fig. 1. Since the direction of this lateral impulse component depends on the target orientation, which is typically unknown, the controllability of the removal endeavor has to be considered.

Fig. 1

Laser-induced momentum during a multiple-pulse engagement of space debris removal. During the flyover of the debris object, N laser pulses are applied from a ground-based high-energy laser station. The induced momentum components Δpax alongside the beam propagation axis are sketched in white, whereas the plane perpendicular to the beam axis, which can be associated with lateral momentum coupling Δplat in unpredictable direction, is highlighted in yellow. Background photo: Paul Wagner, DLR.

OE_58_1_011004_f001.png

Therefore, our concern is whether those unpredictable lateral momentum components can be neglected and accordingly whether they scatter in magnitude and orientation of the thrust angle α poses a serious challenge for the reliability and safety of laser-based debris removal operations.

Moreover, recent studies have shown that the residual heat stemming from the laser ablation process is likely to rapidly heat up the debris target during repetitively pulsed laser irradiation.14 Therefore, irradiation restrictions might apply in order to avoid target melting and compaction, which would imply the risk of losing track of the target. Hence, our studies aim to assess this secondary operational risk in laser-based removal of space debris as well.

1.4.

Scope of Work

The paper is organized as follows: The underlying physics of laser ablation is depicted in Sec. 2.1, where both momentum generation as well as the generation of residual heat at the target are described. For the quantification of these processes, hydrodynamic (HD) simulations have been carried out, which are explained in greater detail in Sec. 3.1. Postprocessing of simulation data yields a database on thermomechanical interaction that allows for laser parameter studies on momentum coupling and residual heat, cf. Sec. 4.

Moreover, the established database serves for configuration of our simulations on laser interaction with irregularly shaped debris targets whose theoretical foundations are outlined in Sec. 2.2. The corresponding numerical implementation of this interaction is given in our raytracing-based code Expedit, which is explained in Sec. 3.2. The code allows for Monte Carlo studies on randomly generated target geometries and yields some insights on the impact of target orientation and hit accuracy, which are shown in Sec. 5.

Although Expedit, in a more simple version, was introduced earlier in our preceding paper12 as a stand-alone program, we now present its implementation as an interaction module in a code that models orbital propagation of space assets. Therefore, some basics on orbital propagation and orbit modification are treated in Sec. 2.3, followed by a short description of the propagator code in Sec. 3.3. In the corresponding result section (Sec. 6), we highlight first findings on single overpass engagements for the scope of space debris removal.

2.

Theoretical Considerations

2.1.

Thermomechanical Coupling in Laser Ablation

2.1.1.

Laser-ablative momentum generation

In the concept studies on laser-based space debris removal, usually rough estimates on laser-ablative momentum coupling can be found: data on optimum impulse coupling cover a range from 14  μN/W for aluminum10 via 75  μN/W for aluminum alloys9 to 160  μN/W for Kevlar,9 and sometimes are generalized covering simply one order of magnitude giving 10 to 100  μN/W for common debris materials.11 The great variance of those data stems from the manifold dependencies of momentum coupling on both target material and laser parameters. Hence, aiming for precise simulation results, these dependencies have to be taken into account.

First of all, the threshold fluence Φ0 for laser ablation has to be considered. Following Ref. 15, Φ0 depends on the laser pulse length τ as follows:

Eq. (1)

Φ0=ΔHvs·Aopt1·(Dvs·τ+αopt1),
where ΔHvs is the effective enthalpy of stationary evaporation, Dvs is the heat diffusivity in the case of stationary evaporation, Aopt is the target absorptivity, and the αopt the optical absorption coefficient of the target. Though laser-induced momentum coupling can already be achieved below the ablation threshold, namely by photon pressure, we restrict our considerations on laser-ablative momentum coupling, which yields coupling coefficients, which are typically three to four magnitudes higher than in the case of pure photon pressure, where coupling is limited to cm=6.7  nN/W.16

For laser-ablative momentum coupling, Phipps introduced an elegant model to describe the dependency of cm on material properties and laser parameters, in particular on the laser fluence.16 Three different regimes of laser-ablative momentum coupling were found with respect to different fluence ranges: at low fluences above the ablation threshold Φ0, momentum coupling in the so-called vaporization regime can be described according to

Eq. (2)

cm,v=2(Toptξ1)(ϱ/α)lnξΦ0·ξ2,
where ϱ is the target density, α is the optical absorption coefficient, ξ=Φ/Φ0 is the normalized fluence, and Topt is the transmissivity.

At higher fluences, interaction of the laser pulse with already ablated material has to be considered as well: vaporized surface material rapidly expands and forms a so-called ablation jet with particle velocities that might exceed several kilometers per second. The jet is aligned to the local surface normal and exhibits a certain divergence angle, which depends on both material and laser parameters.17,18 Hence, a fraction of the incoming laser light can be absorbed by the jet during the further temporal course of the laser pulse leading to an increase of heat and pressure inside the jet.

Above the threshold Φp=τ×4.79×104  J/cm2, plasma ignition occurs in the ablation jet. The ionization fraction ηi, given by ηi=ni/(n0+ni), where n0 and ni are the number densities of neutrals and ions, respectively, increases with the fluence yielding increasing plasma shielding of the target surface, which reduces momentum coupling. Finally, when the fluence is sufficient for a fully ionized jet, ηi=1, this transition regime is followed by the plasma regime, where momentum coupling can be described according to

Eq. (3)

cm,p{A/2[Z2(Z+1)]1/3}9/16/A1/8(Iλτ)1/4,
where A is the average atomic mass, Z=ne/ni is the mean ionization state of the plume, and I=Φ/τ is the average intensity of the laser pulse.

In the generalized model proposed in Ref. 16, the findings on momentum coupling for the vaporization regime and the plasma regime are joined using an interpolation, which is governed by the ionization degree of the jet:

Eq. (4)

cm(Φ)=[1ηi(Φ)]·cm,v(Φ)+ηi(Φ)·cm,p(Φ).

This model holds for laser pulses >10 to 100 ps. For ultrashort pulses, e.g., for photomechanical ablation processes like spallation, however, the here applied assumptions on ablation processes fail, since jet formation commences when the laser pulse is already over. Nevertheless, a similar behavior of cm(Φ) can be found.19

2.1.2.

Residual heat in laser ablation

Although the laser-induced sudden and strong increase of heat and pressure at the target surface yields vaporization and/or spallation of a part of the surface, the heat affected zone of the target surface may comprise much more material than the ablation jet itself. It can be seen from the simulation results shown in Fig. 2(a) that for the short-pulse regime the remaining target surface layers are rapidly heated to 2000 K and beyond. In contrast, rapid spallation of a comparatively large amount of surface material with ultrashort pulses yields lower initial temperatures at the surface of the remaining target. Although in the short-pulse regime ablation is mainly governed by surface vaporization of the heated material, fast heating with an ultrashort laser pulse raises a strong shock wave followed by a large rarefaction wave, which exceeds the maximum tensile strength of the material.19 It can be seen from the discontinuities in Fig. 2(b) that material spallation occurs several times at the surface before the rarefaction wave sufficiently relaxes. Nevertheless, though frequently referred to as cold ablation, a significant amount of heat remains in the target after spallation, not to forget the secondary contribution by the thermalization of the laser-induced shock wave inside the material.

Fig. 2

Heat distribution in laser ablation with: (a) short and (b) ultrashort laser pulses. Laser parameters: λ=1064  nm, ϑ=0  deg, (a): τ=50  ps, Φ=1.49  J/cm2, (b): τ=5  ps, Φ=0.74  J/cm2, results from 1-D HD simulations with Polly-2T. The target is 10- to 50-μm-thick and its surface is located at x=0. The laser pulse exhibits a Gaussian temporal shape, which peaks at t=0. Polly-2T is described in greater detail in Sec. 3.1.

OE_58_1_011004_f002.png

The overall amount of heat Qres after ablation is usually quantified by the coefficient of residual heat ηres using the ratio:

Eq. (5)

ηres=Qres/EL.
Typical experimental data on ηres are 15% to 25% throughout a pulse length range from 60 fs to 6  μs, but values of up to 40% and down to 10% can be found as well.20,21 Since this heat load per pulse contains much more energy than that expected to be reradiated from a debris during repetitive laser irradiation, heat accumulation at the target has to be considered—an effect that is already reported as a problematic issue in laser material processing,22 where even more fortunate circumstances are usually present (convection cooling, heat sink, ultrashort laser pulses, etc.).

To account for this problem, we have introduced14 the thermomechanical coupling coefficient ctm defined by

Eq. (6)

ctm(Φ)=Δp/Qres=cm(Φ)/ηres(Φ),
describing the laser-induced momentum that can be achieved, when a certain amount of residual heat can be taken into account for. Then the maximum velocity increment that can be achieved without target melting can be assessed using:

Eq. (7)

Δvmax=cm(Φ)ηres(Φ)/cp(TmT1),
where cp is the specific heat of the target, T1 is the initial temperature before the first-laser pulse, and Tm is the melting point.

2.2.

Thermomechanical Coupling with Irregularly Shaped Targets

Although in the previous section, laser–matter interaction has been described in detail with respect to the laser parameters Φ, τ, and λ and material, the specific target shape has to be addressed as well since the target geometry significantly affects momentum magnitude and, in particular, its direction. A first approach has been undertaken in this regard by Liedahl et al.23 considering that each surface element of the target exhibits a momentum component that is aligned alongside the local surface normal. For this purpose, the area-matrix concept was introduced enabling the analytical calculation of laser-imparted momentum on simple target geometries. As a drawback, however, the fluence dependency of cm had to be neglected in that method. Therefore, we proposed an extension of the area-matrix concept taking into account for fluence variations throughout the target surface, which now requires a numerical treatment using:

Eq. (8)

pj(r)=cm(ΦL,ϑ)·ΦL(r)·dAj(r)·cosϑj(r)dn^j(r),
where pj(r) denotes the momentum induced at the position r on the j’th irradiated surface element dAj with the local surface normal n^j.12 ϑj is the local incidence angle of the incoming laser beam, given by a fluence distribution ΦL(r). Correspondingly, this approach can be extended for residual heat using:

Eq. (9)

dQj(r)=ηres(ΦL,ϑ)·ΦL(r)·dAj(r)·cosϑj(r).

Overall momentum coupling to the target as well as acquired heat can be obtained easily by summation over all nonshadowed surface elements, p=jpj and Qres=jdQj, respectively. For comparison with momentum coupling in one-dimensional (1-D) calculations, a combined efficiency ηc was introduced by Phipps et al.9 taking into account for the effects of “improper thrust direction on the target, target shape effects, tumbling, and so on:”

Eq. (10)

Δvax=ηc·cmΦL/μ,
where Δvax is the velocity increment alongside the laser propagation axis and μ is the target areal mass density assessing ηc30%.

The target areal mass density is given by μ=m/Ax, where m is the target mass and Ax is the target cross-sectional area. Though not clearly stated in Ref. 9, it should be noted that Ax is usually not identical with the cross-sectional area of the target with the laser beam but defined by the characteristic length Lc of the target using24 Ax0.56·Lc2 for fragmentation debris with a characteristic length exceeding 1.7 mm, following the NASA standard breakup model (SBM). Lc, in turn, is given by the arithmetic mean of the characteristic dimensions X, Y, and Z, which are the orthogonal set of the respective maximum target extensions.5

2.3.

Orbit Modification by Laser-Induced Momentum

Applying a velocity increment Δv to an orbital asset yields a modification of its trajectory. A detailed analytical treatment is given in Ref. 25 showing that, although an instantaneous altitude increase might occur, the orbital eccentricity is raised such that the target’s perigee is decreased. Typically, a Δv of 150 m/s is reported to be sufficient for perigee lowering from low Earth orbit (LEO) down to xp=200  km causing debris removal by burn-up in the upper atmosphere.25

With respect to multipulse irradiation of irregularly shaped debris targets, lateral impulse components have to be considered, cf. Fig. 1. In our previous work, we have shown that the irregularity of the target shape easily yields laser-induced rotation, which, due to the typically missing synchronicity of rotation and pulsed laser irradiation, leads into a chaotically spinning behavior of the object.12 Given an arbitrary target orientation, however, evenly distributed lateral momentum components in arbitrary directions might nearly average out for a great number of laser pulses during the removal engagement.

With this simplification, the findings on average axial momentum coupling pax can be employed for a reliable prediction of the overall imparted momentum Δp and hence the modified debris trajectory using

Eq. (11)

Δpi=1Npax,ipax×i=1Nn^i,
where n^i is the unit vector pointing in the laser propagation direction at the ith laser pulse. The simple product in Eq. 11 suggests that two important fields of laser-based debris removal can be optimized independently: First, laser–matter interaction, given by pax, and second, the strategy of pulsed debris irradiation as represented by i=1Nn^i. Although interaction is determined by the laser configuration, the irradiation strategy can be considered independently with respect to the orbital parameters, e.g., maximum elevation βmax from the laser station, orbit altitude z and orbital eccentricity ε, and the suitable engagement interval [βin,βout].

Thanks to laser-based debris monitoring techniques, which already exhibit a great precision in orbit determination,26 the term related to the irradiation strategy in Eq. 11 is accessible by the optical measurement of both azimuthal angle φ and elevation angle β during the overfly, i.e., n^(t)=[cosβ(t)·cosφ(t),cosβ(t)·sinφ(t),sinβ(t)] can be directly derived from laser-based measurements.

3.

Methods

3.1.

Hydrodynamic Simulations on Thermomechanical Coupling

In order to derive the dependencies of thermomechanical coupling from the laser pulse parameters, HD simulations have been carried out with the 1-D code Polly-2T, provided from M. Povarnitsyn at the Joint Institute for High Temperatures at the Russian Academy of Sciences, Moscow. In Polly-2T, the target material is simulated using semiempirical equations of state for ion lattice and electron gas taking also into account for metastable states. Laser-beam coupling into the target is modeled with the Helmholtz equation. Heat distribution within the material is calculated using the two-temperature model. Interaction of the laser radiation with the induced material response is in particular accounted for by the usage of dynamic models for dielectric permittivity, heat conductivity in the electron gas and electron–phonon coupling, which cover a great range of electron temperatures.27 A detailed description of the HD code is given in Ref. 28.

Postprocessing of HD simulation results comprises the calculation of imparted momentum in laser ablation yielding cm for the chosen laser parameter set of Φ,τ,λ,ϑ and polarization. A corresponding laser parameter study with corresponding details is described in Ref. 19.

For the scope of our study, the calculation of residual heat was added to simulation postprocessing. Therefore, the temporal course of additional thermal energy after ablation was derived and the course of additional kinetic energy was analyzed with respect to imparted recoil and kinetic energy associated with shock wave formation. The latter fraction was added to the increase of thermal energy since thermal relaxation of the shock wave can be assumed. An approximation was undertaken to estimate the boundary value of residual heat from the ablation event. The results of the simulations are analyzed in Sec. 4.

3.2.

Simulation of Laser Interaction with Irregularly Shaped Targets

Expedit is a code written in C++ for the calculation of laser–matter interaction with arbitrarily shaped targets. Targets are given as sets of finite surface elements and interaction with the discretized laser beam is calculated for each ray-surface intersection, provided no self-shadowing occurs. This method enables to attribute a specific fluence Φ(r) to each ray and, hence, to derive a fluence-specific imparted momentum Δp[Φ(r)] to each surface element, cf. Eq. 8. In this regard, fit functions describing cm(Φ) are taken from Polly-2T results and used for configuration of Expedit. The initial version of Expedit is described in greater detail in Ref. 12, an experimental validation can be found in Ref. 13.

Expedit was recently rewritten and upgraded during the seconds author’s thesis.29 This upgrade comprises the implementation of Expedit on a graphics processing unit using Cuda and, for raytracing operations, NVidia Optix. Hence, the code is massively parallelized using 3840 cores (NVidia Quadro P6000), which yields a great performance increase compared to Ref. 12. Moreover, a discretized calculation of residual heat following Eq. 9 was implemented as well and code-wrapping allows user-friendly employment via a Python scripting. The results of our recent studies with the Expedit module are reported in Sec. 5.

3.3.

Orbital Propagation of Space Objects

For calculation of the orbital propagation of the space debris target, a Python script was written using a Newtonian approach, i.e., taking only into account for the Earth’s gravitational field but neglecting secondary factors such as atmospheric drag, radiation pressure, solar, and lunar gravitation. The propagator is initialized with the target’s position x and velocity v at a certain point in time. Then x and v are propagated considering gravitational acceleration via the velocity verlet algorithm with a time step size of Δt=100  ms as described in detail in Ref. 29.

Propagator initialization can be done by choosing a circular orbit at a certain altitude over a specific location on Earth. Moreover, it is possible as well to use orbital datasets in the two-line-element format. Additionally, the target’s orientation is specified with propagator initialization. Target rotation can be propagated as well, but this feature is not used in the work presented here. Instead, random orientation of the target before each laser pulse is chosen.

The propagator calls Expedit as a laser–matter interaction module each time a laser pulse is initiated. The Expedit module returns the corresponding increments of translational and rotational momentum Δp and ΔL, respectively, which are considered for further orbital propagation of the target. The respective simulation results on multipulse irradiation of space debris are shown in Sec. 6.

4.

Laser–Matter Interaction Database

In general, solid-state lasers with short pulses are proposed for debris removal. It can be seen from Table 2 that in the case of ground-based debris removal, the range of recommended laser parameters is rather fixed. The laser wavelength is in the visible or near-infrared part of the spectrum, which can be ascribed to both atmospheric transmittance as well as the availability of suitable high-energy laser sources. In contrast, for space-based laser operation, shorter wavelengths are easily possible.

Table 2

Literature suggestions on laser configurations for removal of space debris in the size range of 1 to 10 cm: pulse energy EL, pulse repetition rate frep, pulse duration τ, laser wavelength λ, and spot diameter ds at the target’s position.

ConceptTypeEL (kJ)frep (Hz)τ (ns)λ (nm)ds (cm)
ORION7Ground-based2014053040
CLEANSPACE8,31Ground-based>10>105 to 50100060
LODR9Ground-based8.5135106031
LODR9Space-borne2330.126023
L’Adroit32Space-borne0.38560.135522
ICAN11Space-borne0.001 to 0.11000 to 90,0001×106 to 110001 to 10

For the pulse length, a similar issue applies. The propagation of short high-energy laser pulses is again limited by the Earth’s atmosphere, namely by air breakdown at high intensities. For space-based operation, much shorter pulse lengths down to femtoseconds are easily conceivable, cf. Table 2. In turn, due to the lower ablation threshold which scales with τ, less pulse energy is needed.

Since aluminum appears to be a rather prevalent material in small-sized space debris,30 it constitutes a good starting point for our considerations on laser–matter interaction. Following the proposals from the literature in Table 2, we performed a laser parameter study, using Polly-2T for laser–matter interaction with aluminum targets. Our study covers the wavelength range of λ=248  nm to 1.536  μm, pulse lengths from τ=100  ps to 10 ns and fluences starting from Φ=0.1  J/cm2 up to 100  J/cm2. A detailed description on their analysis is given in Ref. 19.

4.1.

Momentum Coupling

Results on momentum coupling obtained from simulation postprocessing are depicted in Fig. 3. cm(Φ) shows the typical behavior of onset of momentum coupling at Φ0, rapid increase of cm in the vaporization regime in junction with a slight decrease in the transition regime until the decrease of cm dominates for higher fluences due to plasma shielding. For the usage in our simulations on debris removal and for the scope of a quantitative analysis on momentum coupling, the fluence dependency of the simulation results was fitted using:13

Eq. (12)

cm(Φ)ΦΦ0ΔΦ+(ΦΦ0)·b·12.46·A7/16·(τλ·Φ)c.
which is valid for ΦΦ0, whereas cm(Φ<Φ0)=0. Note that in Eq. 12 Φ is given in J/cm2. For wavelength λ and pulse length τ, however, the usual SI units are used (m and s, respectively).

Fig. 3

Momentum coupling with aluminum targets at λ=1064  nm for various laser pulse lengths. Results from 1-D HD simulations with Polly-2T.

OE_58_1_011004_f003.png

The fit function in Eq. 12 exhibits four free parameters Φ0,ΔΦ,b,c and was deduced from Eq. 4 neglecting the vaporization regime since in our simulations the thresholds for material ablation Φ0 and plasma ignition Φp were found to be rather similar. Assuming for simplification that ηiΦ, the transition regime can be characterized by the transition fluence Φt, where half of the jet is ionized, ηi=0.5, which occurs at Φt=Φ0+ΔΦ. In practical terms, the transition range ΔΦ is more or less related to the position Φopt of the curve maximum, whereas the peak parameter b describes the maximum value of momentum coupling cm,opt, and the plasma exponent c gives the negative slope of cm in a log–log plot against the fluence Φ.

Selected results on fit parameters Φ0,ΔΦ,b,c are shown in Table 3. These data allow for an assessment of laser parameter optimization in space debris removal: with respect to the minimum required fluence, Φ0 gives an estimate for the laser pulse energy needed at a given laser spot size at the debris position. As it can be expected from Eq. 1, Φ0 increases with τ, furthermore, for ultrashort pulses, i.e., below τ=50  ps in the case of aluminum,19 Φ0 stays more or less constantly on a rather low value.

Table 3

Fit parameters and corrected regression coefficient R2 for momentum coupling, results from HD simulations with Polly-2T on laser ablation of aluminum at λ=1064  nm. For the atomic mass, A=26.98 was used.

τ (ns)Φ0 (J/cm2)ΔΦ (J/cm2)b (N/MW)cR2
1×1040.37±0.011.03±0.241.696±0.2990.304±0.0450.981
1×1030.26±0.014.61±1.803.267±1.0980.621±0.0850.979
0.010.27±0.012.24±1.281.145±0.3100.793±0.1340.970
0.10.53±0.010.39±0.060.343±0.0060.344±0.0200.980
11.07±0.031.43±0.370.266±0.0100.474±0.0510.959
103.03±0.084.92±1.960.195±0.0150.583±0.1020.953

For high fluences, it can be seen from the plasma exponent c that the strength of plasma shielding is significantly pronounced with greater pulse lengths, which is likely due to the longer interaction time of the expanding plume with the ongoing laser pulse. The high data for c in the picosecond range, however, are somewhat misleading here since in the absence of plasma shielding for ultrashort pulses cm is limited by the maximum tensile strength of the material.19 Overall, the plasma exponent c is considerably higher than the theoretical value of 1/4 derived from Phipps, cf. Eq. 3. This can be ascribed to the simplification realized in our fitting function, Eq. 12, which neglects the fluence dependency of the mean ionization state Z=Z(Φ) in Eq. 3.

The fluence Φopt for optimum impulse coupling can be obtained using:

Eq. (13)

Φopt=Φ0+(1/c1)ΔΦ/2+q,
with the parameter q defined as

Eq. (14)

q=ΔΦ[Φ0/c+ΔΦ/(4c2)+ΔΦ/4ΔΦ/(2c)].

Then cm,opt can be derived using Eq. 12, cf. Fig. 4. Simulation results show a clear increase of momentum coupling with decreasing wavelength, which corresponds to the increasing absorptivity in this case. Basically, this trend is confirmed by experimental data where ablation by visible laser light yields approximately twice as much momentum or even more than in the near-infrared. A specific dependency of cm,opt on the pulse length is not clearly pronounced.

Fig. 4

Optimum laser-ablative momentum coupling of aluminum, results from 1-D HD simulations with Polly-2T in comparison with experimental data. References for experimental data: a: Ref. 33, b: Ref. 34, and c: Ref. 35, respectively.

OE_58_1_011004_f004.png

Summing up, in terms of momentum coupling, short-laser pulses at a short wavelength seem to be advisable from a theoretical point of view. Nevertheless, the pulse length specific technological challenge to achieve the desired pulse energy has to be taken into account as well as, if a ground-based technology is chosen, high-intensity pulse propagation through the atmosphere.

4.2.

Thermal Coupling

Although in the past laser parameter studies often focused on the maximization of cm as the core figure of merit in laser-propulsive issues, recent considerations on the accumulation of residual heat from laser ablation gave rise to the usage of an alternative metric for laser parameter optimization.14

Similar to the treatment of our cm data from HD simulations, the results on the residual heat ηres in laser ablation, which are depicted elsewhere35 were approximated using the following empirical function:

Eq. (15)

ηres(Φ)=a0+a1Φ+a2Φ21+a3Φ+a4Φ2+a5Φ3(Φ>0)
yielding the fit parameters a05 depicted in Table 4. The enumerator of Eq. 15 describes, roughly speaking, the course of ηres(Φ) in the regime of heating and melting below Φ0, whereas the denominator more or less takes into account for ηres(Φ) in the ablation regime. For short-laser pulses ηres amounts to 20% and more around the onset of ablation and shows a slow but continuous decrease with increasing fluence. In contrast, ηres does not exceed 7% for ablation with ultrashort pulses. Below the ablation threshold, ηres equals more or less its peak value in the case of ultrashort pulses, whereas for short pulses a continuous increase of ηres with increasing fluence can be found.

Table 4

Fit parameters for residual heat ηres(Φ) at λ=1064  nm, results from HD simulations with Polly-2T on laser ablation of aluminum.

τ (ns)a0a1 (J−1 cm2)a2 (J−2 cm4)a3 (J−1 cm2)a4 (J−2 cm4)a5 (J−3 cm6)R2
1×1040.072±0.0050.09±0.020.036±0.0111.12±0.340.48±0.200.013±0.0060.696
1×1030.071±0.0030.09±0.010.035±0.0071.18±0.220.46±0.140.018±0.0050.844
0.010.058±0.0050.12±0.020.070±0.0241.74±0.250.82±0.250.136±0.0570.936
0.10.067±0.0040.16±0.030.320±0.0482.38±0.122.11±0.130.607±0.1220.997
10.067±0.0050.09±0.020.055±0.0121.15±0.050.35±0.030.059±0.0210.988
100.063±0.0030.03±0.010.010±0.0040.46±0.020.07±0.010.003±0.0030.999

Albeit restricted on results from simulations and aluminum as a target material, the fit parameter data from Tables 3 and 4 allow to depict the usage of the thermomechanical coupling coefficient ctm=cm/ηres as a new metrics for laser parameter optimization in laser-ablative propulsion issues. It can be seen from Fig. 5 that ultrashort laser pulses clearly outperform short pulses, i.e., ultrashort pulses are more likely than longer pulses to achieve the required velocity increment Δv in debris removal without heating the target too much. Their superior performance in thermomechanical coupling compared to laser pulses of 100 ps and longer can mainly be attributed to their significantly lower thermal coupling cf. Sec. 2.1.2 and Ref. 35.

Fig. 5

Dependency of thermomechanical coupling from laser parameters. Results from HD simulations with Polly-2T on laser ablation of aluminum at λ=1064  nm wavelength, circular polarization, and perpendicular beam incidence. Note that solid lines do not represent fitting curves of ctm(Φ) here but have merely derived from Eq. 6 using the tabulated fit parameters for Eq. 12 and Eq. 15, respectively.

OE_58_1_011004_f005.png

In contrast, for the short-pulse regime, the strong limitation in Δv without target cooldown is mirrored at the secondary y-axis of Fig. 5, cf. Eq. 7. In particular, debris removal in a single-laser station overpass does not seem to be feasible since the required velocity increment of ΔvLEO=150  m/s can hardly be achieved without intermediate target cooldown for laser pulses in the nanosecond range. This contradicts earlier removal concepts within a single transit using high-power laser systems in the nanosecond range, cf. Table 2. Nevertheless, validation of our findings by experimental data for various relevant debris materials is needed.

5.

Interaction with Space Debris Targets

As a baseline setup for our studies on interaction with debris targets, we chose a laser configuration proposed in Ref. 8, which is a ground-based high-energy laser with EL=25  kJ, τ=10  ns, and λ=1064  nm. With ds=2/3  m, referring to the Φmax/e2 of the Gaussian beam profile, an average fluence Φ=7.2  J/cm2 can be achieved at the target position.

As simulation targets, with the lack of available experimental or real-world data, we randomly generated 100 ellipsoids with characteristic dimensions X,Y,Z under the constraints X/Y[1,2], Y/Z[1.5,60.3], Z[0.3  mm,5  cm], Ax/m[0.1  m2/kg,0.2  m2/kg], and Lc[0.01  m,0.1  m]. These ratios are characteristic for debris fragments from ground-based satellite crash tests reported in Refs. 5 and 36, respectively, and their consideration allows for an approximate calculation of laser-target cross-sectional area and momentum directionality in our studies. However, the ellipsoid shape is still a simplified geometry and does not take into account for the potential multitude of debris shapes that might be found in this size range. This issue has to be subject to future studies.

Although Ax/m depends on Lc in the SBM, we have used equally distributed random functions for the sake of simplicity, which is justified by the large scatter in the SBM. For target generation, a Python script was used initiating target generation with random values for X/Y and Y/Z and Z, respectively, under the above-mentioned constraints. If the corresponding ellipsoid specified by X, Y, and Z matched the given constraints for Ax/m and Lc, the geometry was selected for target generation resulting in a Wavefront.obj file and attribution of the material density to the target. For raytracing, a spatial resolution of 0.1 mm was used.

5.1.

Target Orientation

For our Monte Carlo simulation, each one of the 100 different targets was placed 2000 times in the laser spot center for a single shot. Each time the target orientation was chosen randomly, whereas the target’s center of mass was aligned to the laser spot center. Mean axial velocity as well as mean thrust angle between laser-beam axis and thrust vector direction were obtained by averaging over all shots for each target, as can be seen from Fig. 6.

Fig. 6

Induced axial velocity increment Δvax (green data plot) as well as thrust angle α (blue plot) versus area-to-mass ratio Ax/m of 100 different flake-like aluminum targets. Each data point represents averaging from 2000 Monte Carlo samples with random orientation. The corresponding statistical errors amount to <4% at 95% confidence level.

OE_58_1_011004_f006.png

The simulation results shown in Fig. 6 underline the validity of Eq. 10 for the axial velocity increment: The more cross-sectional area is exposed to the laser beam at a given target mass, the higher is the acquired velocity increment. In particular, it is shown here that the common definition of Ax for space debris, cf. Sec. 2.2, gives a reasonable average cross section in the case of laser irradiation as well.

In contrast, α is almost independent of Ax/m. It amounts to 45  deg±18  deg in our simulations, which implies a mean loss of efficiency in Δvax by a factor of 11/2 due to random target orientation. Lateral Δv components, however, have an unpredictable direction in the plane perpendicular to the propagation axis and might cancel out more or less during a multipulse engagement, provided the target spins rather fast.

The large scatter in α corresponds to the large scatter in Δvax depicted in Fig. 6, which is even more pronounced due to the fluence dependency of cm with cm(Φ)=cm(ΦLcosα). Hence, the relative error σΔvax/Δvax in Δvax due to the unknown target orientation amounts to 86.3%±2.9%. Moreover, the relative error of Δvax due to the scatter in the target’s Ax/m throughout the target ensemble is 21.4%, which is nearly identical to the scatter of Ax/m itself, cf. Eq. 10.

For laser-ablative heating a relation similar to Eq. 10 applies:14

Eq. (16)

ΔT=cp1·ηresΦL/μ,
which is depicted in Fig. 7. Again, the greater the area exposed to the laser beam is at a given mass, the higher is the laser-induced temperature increase. The practical implication of the findings from our simulations is that the precise assessment of the target’s temperature increase is strongly impeded due to missing knowledge of its orientation as well as possible uncertainties in the determination of its mass areal density μ. Hence, large safety margins have to be applied for the maximum number Nmax of laser pulses during one transit in order to avoid target melting, cf. Fig. 7.

Fig. 7

Induced temperature increment ΔT per pulse versus area-to-mass ratio Ax/m of 100 different flake-like aluminum targets. The corresponding maximum laser pulse number Nmax=(TmT1)/ΔT to avoid target melting is shown as a secondary y-axis(T1=0).

OE_58_1_011004_f007.png

5.2.

Laser Pointing Accuracy

As a next step, we introduced errors in spot positioning in our simulations. For this purpose, we varied the position of the object’s center of mass for each pulse within the laser spot using a random normal distribution N(0,σh), where σh is the hit uncertainty of the laser. For this uncertainty, we considered three parameters in a form σh=σt2+σp2+σs2, where σt is the tracking accuracy, σp is the pointing accuracy, and σs is the uncertainty due to beam wander. In Ref. 8, σt=0.1  μrad and σp=0.07  μrad was stated as a requirement for laser-based debris removal. In contrast, σs was assessed to be significantly higher in Ref. 29 amounting to roughly 0.4  μrad. Hence, we have marked the resulting value of σh=0.42  μrad as orientation in our simulation result, cf. Fig. 8. The corresponding lateral uncertainty in spot positioning at a distance of 800 km (altitude of an SSO) is given there as the primary x-axis.

Fig. 8

Combined efficiency ηc and relative standard deviation σΔvax/Δvax of the axial velocity increment: dependency on hit uncertainty.

OE_58_1_011004_f008.png

For the assessment of our simulation results, we calculated the combined efficiency ηc according to Eq. 10, which is depicted in Fig. 8. It should be noted here that deviating from Phipps’ definition in Ref. 9, we now take into account for misalignment and the spatial distribution of the beam profile, which is both not considered there. In particular, we set Φ=Φ, which in terms of the interaction process yields a overestimation of ηc if the target is located in the vicinity of the beam center, where Φ(0)=2Φ and correspondingly an underestimation for a placement at the outer rim of the spot. Nevertheless, ηc remains a useful figure of merit in terms of technology assessment where a certain σh can be specified for a given spot size, distance, and pulse energy.

In the case of a very low-hit uncertainty up to 10  cm corresponding to the uncertainty contributions considered in Ref. 8, σh=σt2+σp2=0.122  μrad, the combined coupling efficiency is rather constant at a level of ηc0.7. Moreover, the relative error of Δvax increases from 86%, cf. the previous section, to 300%, which implies a relative uncertainty due to positioning errors of up to 290%.

For greater hit uncertainties, which are likely when effects like uncompensated beam wander are considered as well, the momentum coupling efficiency decreases linearly with increasing σh, whereas the relative error in Δvax increases proportionally to σh. This strong performance loss in ablative momentum generation can be ascribed to the lower fluence at the outer parts of the laser spot as well as the decreasing hit rate for greater positioning uncertainties. Hence, an overall hit accuracy of 0.1  μrad seems to be a reasonable system requirement for the selected laser spot size.

Concerning current laser technology, pointing accuracies down to a standard deviation of 2.6  μrad have been reported37 and currently a guide star system for turbulence compensation is designed for a residual on-sky jitter of 500 nrad rms.38 Admittedly, these specifications do not yet match the requirements from our simulations. However, promising new technologies like Brillouin-enhanced four-wave mixing for turbulence compensation25 and dynamic mode excitation for the control of mode instabilities39 might yield a significant improvement of pointing stability for future ground-based laser systems. In general, however, laser system specification exhibits a greater complexity: pointing requirements would be more relaxed if a greater laser spot can be chosen. This, in turn, would require higher pulse energies in order to maintain the required fluence. In this regard, laser-beam quality, the Strehl ratio of the transmitter as well as the capability of adaptive optics to compensate beam broadening define the technological boundary conditions of in-orbit laser–matter interaction.

6.

Orbit Modification under Repetitive Laser Irradiation

Though the scatter in laser-induced momentum transfer, as described in the previous section, is rather large, the overall impact to the trajectory of a space debris target has to be quantified separately, since |Δp||p|, which implies that in general the whole laser-debris engagement might still exhibit a satisfying reliability and controllability, i.e., safety.

Therefore, we extended the above-mentioned Monte Carlo studies attributing the debris targets to a specific trajectory passing a ground-based laser removal station taking into account the repetition rates frep and ranges of elevation β proposed in the literature.8

For the purpose of this study, we now restrain to a very simple target, namely a 20×20×1  mm3 aluminum plate, which has already been subject to true-scale debris removal experiments for the validation of Expedit.13 Although the employed laser in the laboratory experiments exhibits a pulse energy of 80  J(λ=1064  nm,τ=10  ns), usage of a reasonable spot size at SSO requires upscaling to EL=20  kJ in our simulations. Note that such upscaled laser systems already exist and have been proposed for space debris removal.40

The target’s orbit altitude was set at 800 km above Earth’s surface with a direct transit above the ground-based laser station. The initial target temperature was set to T0=291.4  K and 398.0 K as to the equilibrium target temperature at the beginning and the end of the sun-illuminated semiorbit, respectively. Target emissivity was set to ε=0.05 according to Ref. 41.

For large distance focusing from the Earth’s surface into SSO, we propose a telescope with DT=8  m diameter and a Strehl ratio of Str=0.4 yielding ds=0.7  m when the laser-beam quality is M2=2. We define the laser irradiation interval from a target elevation of βin=30  deg above the horizon until, shortly after station overfly, βout=100  deg is reached. We assume that the laser keeps track of the target during laser-induced orbit modification, however, with a pointing accuracy of σp=0.42  μrad, cf. Sec. 5.2. Various pulse repetition rates are tested, ranging from 0.1 to 50 Hz. For each irradiation configuration Ns=800 modified trajectories with random target orientation are taken as Monte Carlo samples in order to derive the overall perigee lowering as well as temperature increase after the repetitive laser irradiation. The corresponding results are shown in Table 5.

Table 5

Predictability of debris orbital parameters after irradiation with Np laser pulses at various repetition rates frep. Averaged results from Monte Carlo simulations for scatter of the target’s position r→f directly after irradiation, perigee and apogee altitude, xp and xa, respectively, period T of revolution and temperature after irradiation during dawn (T1) and dusk (T2), respectively, initial orbit altitude: 800 km (circular).

frep (Hz)Npσr→f (m3)xp (km)xa (km)T (s)T1 (K)T2 (K)
0.11836×46×59796.9±1.8800.6±0.56041.8±1.2338±11439±10
0.23550×65×85794.1±2.7801.0±0.66040.3±1.7375±15475±14
0.58880×102×133785.4±3.9802.4±0.96035.7±2.6487±22580±21
1175115×147×185771.2±5.6804.6±1.26028.1±3.7658±29739±26
2351n.d.n.d.n.d.n.d.>Tm>Tm

With respect to the uncertainty of the resulting orbit parameters, it should be noted that the target orientation might have a significant effect on the efficiency of orbit modification. This is obvious when rather flat targets are assumed since the imparted momentum of a single pulse is oriented alongside the surface normal. Therefore, when the target rotation is taken into account within the simulations, the resulting perigee alterations exhibit a comparatively large scatter since the interaction of target rotation rate and pulse repetition rate might yield pulse trains with either frequent fortunate target orientations or with many detrimental target orientations.29 In this regard, the uncertainty derived in our simulations, where the target orientation was randomly chosen for each laser pulse yields a lower scatter of orbital parameters, which might only serve as a rough estimation for the occurring uncertainties. In turn, the efficiency of a removal engagement might greatly benefit from target rotation analysis using sunlight reflectance, cf. Ref. 42. Nevertheless, the uncertainties in prediction of orbital parameters, which are associated with the scatter of our simulation results, cf. Table 5, underline the necessity of a sound covariance analysis of conceivable trajectory modifications. A corresponding collision analysis before laser operation is recommended.

It should be noted here that for LEO objects ESA requires an accuracy of orbital data yielding a 1-sigma position error <40  m×200  m×100  m for at least 48 h after data generation.43 This accuracy requirement, in comparison with the uncertainty of rf, xp, and xa shown in Table 5, underlines the necessity of thorough postirradiation tracking of the debris particle.

With respect to thermal coupling, the results shown in Table 5 suggest that for flat targets a repetition rate of 1 Hz would be the upper limit for irradiation during one transit. For this laser configuration, however, the corresponding perigee lowering at 1 Hz is only Δxp=28.8±5.6  km. This perigee lowering is much too small to achieve an orbit modification yielding xp200  km, which is required for target burn-up. Therefore, the main implication of thermal restrictions is that with this laser configuration it is only possible to remove debris in a series of multiple subsequent overpasses with laser irradiation and intermediate target cooldown, in contrast to the single-pass option presented for such a system as in Ref. 8.

Taking into account for pulse number restriction, it might be rather advantageous in terms of perigee lowering to choose a shorter irradiation period with a subsequently higher repetition rate. However, low-repetition rates are beneficial with respect to intraburst target cooldown, target tracking and, which is more, the required average laser power, which is considerably lower than proposed in the recent studies, cf. Table 2.

7.

Conclusions and Outlook

Applying directed energy to space assets has got several implications on the motion and structural integrity of the irradiated target. Two examples, scatter of imparted momentum and accumulation of residual heat from ablation, have been studied in this paper. In this regard, our findings from simulations clarify that space debris removal by laser ablation is significantly different from simply shooting for debris with a laser like it is sometimes put forth in nonscientific media. Quite the contrary, repetitive laser irradiation during one overfly has to be limited for two reasons related to operational safety.

First, the applied momentum after laser irradiation is associated with a large uncertainty that easily might exceed its predicted average value. Sources of uncertainty are precision limitations in both pointing as well as in the determination of the target’s mass areal density, possibly associated with missing knowledge of the target’s material, its actual orientation and, therefore, the direction of laser-induced thrust. Though random momentum components lateral to the laser-beam propagation axis might cancel out in the long run, the orbital parameters of the debris’ trajectory after irradiation can only be predicted with a remarkable uncertainty, which requires a comprehensive collision analysis in advance to laser operation.

Second, residual heat in laser ablation limits the number of possible pulses during one transit. Depending on the target material, which should remotely be reconnoitered, reasonable safety margins for laser operation should be sought in order to avoid target melting, which would turn debris removal operations into a debris compactor. Beneficially, these thermal issues suggest to operate with an average laser power that is considerably lower than known from earlier concepts.

Though under these boundary conditions high-energy lasers might not appear as a safe tool for delicate in-space operations, it does not require much more than reasonable diligence and the adequate precision of a typical scientific challenge to softly remove debris from orbit. However, in contrast to our previous paper where we compared debris removal with an orbital sweeping broom,12 the findings from this study suggest that removal rather resembles medical surgery using a sharp scalpel.

Disclosures

We declare to have no conflicts of interest related to this work.

Acknowledgments

The authors gratefully acknowledge the scientific contribution of Claude Phipps, Jascha Wilken, Jens Rodmann, Michael Zwilich, and Daniel Burandt in inspiring discussions. We especially thank Mikhail Povarnitsyn for providing us with the Polly-2T code.

References

1. 

D. J. Kessler and B. G. Cour-Palais, “Collision frequency of artificial satellites: the creation of a debris belt,” J. Geophys. Res. Space Phys., 83 (A6), 2637 –2646 (1978). https://doi.org/10.1029/JA083iA06p02637 Google Scholar

2. 

C. Wiedemann and E. Stoll, “Space debris—trends and challenges,” in Presentation at DLR Stuttgart, (2017). Google Scholar

3. 

P. H. Krisko et al., “ORDEM 3.0 and MASTER-2009 modeled debris population comparison,” Acta Astronaut., 113 204 –211 (2015). https://doi.org/10.1016/j.actaastro.2015.03.024 AASTCF 0094-5765 Google Scholar

4. 

S. Flegel, “Maintenance of the ESA MASTER model,” (2011). Google Scholar

5. 

P. Krisko, M. Horstman and M. Fudge, “SOCIT4 collisional-breakup test data analysis: with shape and materials characterization,” Adv. Space Res., 41 (7), 1138 –1146 (2008). https://doi.org/10.1016/j.asr.2007.10.023 ASRSDW 0273-1177 Google Scholar

6. 

W. Schall, “Orbital debris removal by laser radiation,” Acta Astronaut., 24 343 –351 (1991). https://doi.org/10.1016/0094-5765(91)90184-7 AASTCF 0094-5765 Google Scholar

7. 

C. R. Phipps et al., “ORION: clearing near-earth space debris using a 20-kw, 530-nm, earth-based, repetitively pulsed laser,” Laser Part. Beams, 14 1 –44 (1996). https://doi.org/10.1017/S0263034600009733 LPBEDA 0263-0346 Google Scholar

8. 

B. Esmiller et al., “Space debris removal by ground-based lasers: main conclusions of the European project CLEANSPACE,” Appl. Opt., 53 (31), I45 –I54 (2014). https://doi.org/10.1364/AO.53.000I45 APOPAI 0003-6935 Google Scholar

9. 

C. Phipps, “A laser-optical system to re-enter or lower low Earth orbit space debris,” Acta Astronaut., 93 418 –429 (2014). https://doi.org/10.1016/j.actaastro.2013.07.031 AASTCF 0094-5765 Google Scholar

10. 

S. Shuangyan, J. Xing and C. Hao, “Cleaning space debris with a space-based laser system,” Chin. J. Aeronaut., 27 (4), 805 –811 (2014). https://doi.org/10.1016/j.cja.2014.05.002 CJAEEZ 1000-9361 Google Scholar

11. 

R. Soulard et al., “ICAN: a novel laser architecture for space debris removal,” Acta Astronaut., 105 192 –200 (2014). https://doi.org/10.1016/j.actaastro.2014.09.004 AASTCF 0094-5765 Google Scholar

12. 

S. Scharring, J. Wilken and H.-A. Eckel, “Laser-based removal of irregularly shaped space debris,” Opt. Eng., 56 (1), 011007 (2017). https://doi.org/10.1117/1.OE.56.1.011007 Google Scholar

13. 

R.-A. Lorbeer et al., “Experimental verification of high energy laser-generated impulse for remote laser control of space debris,” Sci. Rep., 8 8453 (2018). https://doi.org/10.1038/s41598-018-26336-1 SRCEC3 2045-2322 Google Scholar

14. 

S. Scharring, R.-A. Lorbeer and H.-A. Eckel, “Heat accumulation in laser-based removal of space debris,” AIAA J., 56 2506 –2508 (2018). https://doi.org/10.2514/1.J056718 AIAJAH 0001-1452 Google Scholar

15. 

D. Baeuerle, Laser Processing and Chemistry, 3rd ed.Springer, Berlin (2000). Google Scholar

16. 

C. Phipps et al., “Review: laser-ablation propulsion,” J. Propul. Power, 26 609 –637 (2010). https://doi.org/10.2514/1.43733 JPPOEL 0748-4658 Google Scholar

17. 

A. V. Pakhomov and D. A. Gregory, “Ablative laser propulsion: an old concept revisited,” AIAA J., 38 (4), 725 –727 (2000). https://doi.org/10.2514/2.1021 AIAJAH 0001-1452 Google Scholar

18. 

D. Ali, M. Z. Butt and M. Khaleeq-ur-Rahman, “Ablation yield and angular distribution of ablated particles from laser-irradiated metals: the most fundamental determining factor,” Appl. Surface Sci., 257 2854 –2860 (2011). https://doi.org/10.1016/j.apsusc.2010.10.080 Google Scholar

19. 

S. Scharring, R.-A. Lorbeer and H.-A. Eckel, “Numerical simulations on laser-ablative micropropulsion with short and ultrashort laser pulses,” Trans. JSASS Aerosp. Technol. Jpn., 14 (ists30), Pb69 –Pb75 (2016). https://doi.org/10.2322/tastj.14.Pb_69 Google Scholar

20. 

M. Autric, “Thermomechanical effects in laser–matter interaction,” Proc. SPIE, 3343 354 –362 (1998). https://doi.org/10.1117/12.321570 PSISDG 0277-786X Google Scholar

21. 

A. Y. Vorobyev et al., “Residual thermal effects in Al following single ns- and fs-laser pulse ablation,” Appl. Phys. A, 82 (2), 357 –362 (2006). https://doi.org/10.1007/s00339-005-3412-0 Google Scholar

22. 

R. Weber et al., “Heat accumulation during pulsed laser materials processing,” Opt. Express, 22 (9), 11312 –11324 (2014). https://doi.org/10.1364/OE.22.011312 OPEXFF 1094-4087 Google Scholar

23. 

D. A. Liedahl et al., “Pulsed laser interactions with space debris: target shape effects,” Adv. Space Res., 52 895 –915 (2013). https://doi.org/10.1016/j.asr.2013.05.019 ASRSDW 0273-1177 Google Scholar

24. 

N. L. Johnson et al., “NASA’s new breakup model of Evolve 4.0,” Adv. Space Res., 28 (9), 1377 –1384 (2001). https://doi.org/10.1016/S0273-1177(01)00423-9 ASRSDW 0273-1177 Google Scholar

25. 

C. R. Phipps et al., “Removing orbital debris with lasers,” Adv. Space Res., 49 1283 –1300 (2012). https://doi.org/10.1016/j.asr.2012.02.003 ASRSDW 0273-1177 Google Scholar

26. 

T. Hasenohr et al., “STAR-C: towards a transportable laser ranging station,” in 68th Int. Astronautical Congress (IAC), (2017). https://elib.dlr.de/116523/ Google Scholar

27. 

M. E. Povarnitsyn et al., “A wide-range model for simulation of pump-probe experiments with metals,” Appl. Surf. Sci., 258 9480 –9483 (2012). https://doi.org/10.1016/j.apsusc.2011.07.017 ASUSEE 0169-4332 Google Scholar

28. 

M. E. Povarnitsyn et al., “Dynamics of thin metal foils irradiated by moderate-contrast high-intensity laser beams,” Phys. Plasmas, 19 023110 (2012). https://doi.org/10.1063/1.3683687 PHPAEN 1070-664X Google Scholar

29. 

L. Eisert, “Numerical simulations on momentum coupling and orbit modification in laser-based debris removal,” (2018). https://elib.dlr.de/119400/ Google Scholar

30. 

J. N. Opiela, “A study of the material density distribution of space debris,” Adv. Space Res., 43 1058 –1064 (2009). https://doi.org/10.1016/j.asr.2008.12.013 ASRSDW 0273-1177 Google Scholar

31. 

B. Esmiller, C. Jacquelard and H.-A. Eckel, “CLEANSPACE—space debris removal by ground based laser—main conclusions,” in High Power Laser Ablation/Beamed Energy Propulsion, Presentation on Conf.-DVD, (2014). Google Scholar

32. 

C. Phipps, “L’ADROIT a spaceborne ultraviolet laser system for space debris clearing,” Acta Astronaut., 104 243 –255 (2014). https://doi.org/10.1016/j.actaastro.2014.08.007 AASTCF 0094-5765 Google Scholar

33. 

B. D’Souza, “Development of impulse measurement techniques for the investigation of transient forces due to laser-induced ablation,” (2007). Google Scholar

34. 

S. Scharring et al., “Low-noise thrust generation by laser-ablative micropropulsion,” in Int. Electric Propulsion Conf., (2015). https://elib.dlr.de/97576/ Google Scholar

35. 

C. R. Phipps et al., “Laser impulse coupling measurements at 400 fs and 80 ps using the LULI facility at 1057 nm wavelength,” J. Appl. Phys., 122 193103 (2017). https://doi.org/10.1063/1.4997196 JAPIAU 0021-8979 Google Scholar

36. 

T. Hanada et al., “Outcome of recent satellite impact experiments,” Adv. Space Res., 44 558 –567 (2009). https://doi.org/10.1016/j.asr.2009.04.016 ASRSDW 0273-1177 Google Scholar

37. 

G. Genoud et al., “Active control of the pointing of a multi-terawatt laser,” Rev. Sci. Instrum., 82 033102 (2011). https://doi.org/10.1063/1.3556438 RSINAK 0034-6748 Google Scholar

38. 

M. Lingham et al., “Adaptive optics tracking and pushing system for space debris manoeuvre,” Proc. SPIE, 10703 107030O (2018). https://doi.org/10.1117/12.2313181 PSISDG 0277-786X Google Scholar

39. 

H.-J. Otto et al., “Controlling mode instabilities by dynamic mode excitation with an acousto-optic deflector,” Opt. Express, 21 17285 –17298 (2013). https://doi.org/10.1364/OE.21.017285 OPEXFF 1094-4087 Google Scholar

40. 

A. M. Rubenchik, A. C. Erlandson and D. Liedahl, “Laser system for space debris cleaning,” AIP Conf. Proc., 1464 448 –455 (2012). https://doi.org/10.1063/1.4739899 APCPCS 0094-243X Google Scholar

41. 

D. E. Gray, American Institute of Physics Handbook, 3rd ed.McGraw-Hill, New York (1972). Google Scholar

42. 

B. D. Pontieu, “Database of photometric periods of artificial satellites,” Adv. Space Res., 19 (2), 229 –232 (1997). https://doi.org/10.1016/S0273-1177(97)00005-7 ASRSDW 0273-1177 Google Scholar

43. 

H. Krag, T. Flohrer and N. Bobrinsky, “ESA’s SST activities and plans for the years 2017–2020,” in Defence Satellites, Conf. Presentation, (2016). Google Scholar

Biography

Stefan Scharring received his diploma degree in physics from the University of Freiburg in 2000 and his doctoral degree in aerospace engineering from the University of Stuttgart in 2013. He works as a senior scientist at the Institute of Technical Physics at the German Aerospace Center (DLR). His current research interests cover the field of laser–matter interactions and their aerospace applications, in particular for space debris monitoring and removal.

Lukas Eisert received his bachelor’s degree in simulation engineering from the University of Stuttgart in 2018. For his bachelor’s thesis, he investigated the orbital alternation due to laser ablation. At the moment, he is enrolled in the astronomy master’s program at the University of Heidelberg. He worked as a research assistant at DLR’s Institute of Technical Physics.

Raoul-Amadeus Lorbeer received his diploma degree in physics from the University of Hanover in 2007 and his doctoral degree from the same university in 2012. He is a research scientist at DLR’s Institute of Technical Physics. His current research interests are focused on laser-ablative thrust generation for aerospace applications, in particular for micropropulsion and space debris removal.

Hans-Albert Eckel studied physics at the University of Kaiserslautern, Germany, where he received his doctoral degree in laser spectroscopy in 1996. He is the head of the “Impact, Protection, and Materials Program” at the DLR. Previously, he was the head of the Studies and Concepts Department at DLR’s Institute of Technical Physics. There, his research focused on high-power laser sources and the assessment of future applications for laser-driven space propulsion.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Stefan Scharring, Lukas Eisert, Raoul-Amadeus Lorbeer, and Hans-Albert Eckel "Momentum predictability and heat accumulation in laser-based space debris removal," Optical Engineering 58(1), 011004 (22 September 2018). https://doi.org/10.1117/1.OE.58.1.011004
Received: 30 June 2018; Accepted: 17 August 2018; Published: 22 September 2018
Lens.org Logo
CITATIONS
Cited by 12 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Pulsed laser operation

Laser ablation

Monte Carlo methods

Optical simulations

Aluminum

Laser irradiation

Atmospheric propagation

Back to Top