Anda di halaman 1dari 24

PROPULSION BY LASER ENERGY TRANSMISSION

(Considerations to Selected Problems)


R.E.Lo* (DFVLR – Institute of Chemical Rocket Propulsion, Hardthausen-Lampoldshausen,
Federal Republic of Germany

Abstract
The rewards of laser propulsion are substantial payload gains for missions with high velocity
increment. Laser-assisted chemical propulsion is an effective means to reduce required laser
powers. Payload gains of this new mode of propulsion can be calculated as functions of ratio
of laser power to chemical power, structural mass fraction and velocity increment. Payload
mass fraction can be optimised for minimum laser power requirement. Deep space missions
require very high laser power and/or very small beam divergence. For near-earth missions, the
simultaneous input of laser and solar-power is worthwhile and can be optimized.

I. Mission Considerations

When a spacecraft receives its propulsive energy by laser- or other energy-transmission from
the outside, a gain over chemical propulsion system can only be obtained, if the effective
exhaust velocity is increased above chemically achievable values. The mass of the energy
transmitted is zero for all practical purposes and the on-board propellant mass would have to
be the same for a given propulsion requirement V and a given exhaust velocity, whether it
contains chemical energy or not. For this reason, all missions requiring larger than chemical
exhaust velocities are the only candidates of real interest for laser-propulsion.
However, many of these missions require also large absolute amounts of power. Typical
examples are
- single-stage-to-orbit missions
- two-way tug missions servicing geostationary orbit
- fast perigee-apogee delivery of payloads.

Therefore, aside of pure laser-propulsion with chemically inert propellants, mixed mode laser-
assisted chemical propulsion should be considered as an effective means to reduce required
laser power. If the take-off mass Mo of a spacecraft is simply split into payload Mn, structural
mass Ms and propellant mass Mp

Mo Mn Ms M p (1)

Mn
(2)
Mo

Ms
(3)
Mo

the basic Ziolkowsky equation

l
V c Ln (4)
a
shows (see Fig. 1) that perigee-apogee missions with a typical V of 4000 m/s and in
particular single-stage-to-orbit missions ( V about 10.000 m/s) suffer from a very steep
decrease of the payload mass fraction with c.

Single-stage-to-orbit propulsion with 1 to 5 % of payload is in the tantalizing region of


marginal feasibility with exhaust velocities of 4500 to 5000 m/s, if structural factors of 0.1
or better can be obtained. This may be the example, where some additional beamed energy
may make all the difference between tiny payload fractions of uncertain feasibility and
commercially interesting percentages. Total exhaust jet power P would thus have to consist of
a chemical and a laser contribution.

PCh and PL :
P PCh PL (5)

Its absolute value depends upon thrust F or massflow rate m which, in turn, are determined
by the chosen vehicle accelaeration g:

P 0.5 m c (6)

F mc (7)

F V
P 0.5 F c (8)
1
2 n

gM n
F gM o (9)
a
gM n V
P (10)
a 1
n

2
PCh 0.5
mc Ch (11)

2
PL 0.5 m c
L
(12)

Due to equation (5), CL is not the difference between necessary exhaust velocity c and the
chemically achievable one CCh, but

CL c 2 CCh
2
(13)

Therefore,

2
CCh
PL 0.5 F (14)
c
and similarly
2
cCh
PCh 0.5 F (15)
c
2
gMn C Ch 1
PCh n (16)
2a V a

2
gMn V CCh 1
PL n (17)
2a 1 V a
n
a

The necessary ratio of laser- to chemical power can thus be expressed as a function of the
desired at a given V for values of ß and CCh as they are achievable:

PL 1 V
1 (18)
PCh 1 C Ch
n

In Fig. 2, the combined sum of payload and structural mass fractions is plotted over the
mission propulsion requirement V in multiples of available chemical exhaust velocity.
Since CCh has to be considered in the range of 3000 to 5000 m/s, the value of V of V /CCh
Is in the range of
- 0.1 for attitude control
- 1.0 for perigee-apogee missions
- 2.0 to 3.0 for single-stage-to-orbit and two-way trips.

For any given structural factor ß, the gain in payload fraction is obviously increasing with
.

For a given it increases dramatically with V /CCh. No significant gain is obtained for low
energy missions. On the other hand, additional laser-power of 10 to 50 % of chemical power
may raise the fraction available for masses other than propellants from 0.1 with zero payload
to feasible values.

This can more clearly be shown if payload gain in terms of % of pure chemical propulsion is
calculated as a function of and V /CCh :

100 a ( )
a% f , V / CCh (19)
a 0
Fig. 3 shows how this gain increases with . It is only for high energy missions that the gain
exceeds 50 to 100 % even for small . Vehicles with heavier structures gain more lighter
ones. For example, payload fraction of = 0.1 single-stage-to-orbit vehicle could be
increased from 3.5 % of Mo by a factor of 4 up to 14.3 % by adding once again the amount of
chemical power with beamed energy. With = 0.2, such a vehicle has no positive payload (
is then – 6.5 %). With = 1, it carries 4.3 % of payload.
Fig. 4 shows as a function of ; for high energy missions payload gain increases steeply
with .

Therefore, it may be concluded that the application of beamed energy is worthwhile only in
missions with large V , with ratios V /CCh above 1. Single-stage-to-orbit missions are
considered as an example in the next chapter, however, very similar results could be obtained
for a two-way tug. Most of the mathematical relations derived hold for any V .

II. Laser-Assisted Single-Stage-to-Orbit Vehicles

Such vehicles fall under two restrictions:


- laser power limited design
- limited design.

The following Table 1 gives the values for a single-stage-to-orbit vehicle with and Mn a
little better than the present space shuttle design (for Table 1, see Page 4).

The following conclusions can be drawn from the results:

- Lifting of 30 t of payload into orbit is not a question of megawatts, as they may be


available within several years, it is a question of gigawatts.

- If, however, gigawatts are available, even a small contribution to chemical power
has dramatic effects on vehicle size: with little more than 5.5 GW it can be reduced
by a factor of more than 21!

- There is a region, where vehicle mass is a very sensitive function of laser power. It is
very much worthwhile to get into this region: in the present example, an increase of
0.26 GW from 5.52 to 5.78 will result in a further vehicle mass reduction by another
factor of almost 3!

- At larger laser powers, their application becomes temperature limited. If 10 00 m/s


are considered as maximum achievable exhaust velocity (corresponding to chamber
temperatures of 6000-7000 K), no more than 5.6 GW can be applied (assuming total
conversion).

- Total power requirement goes through a minimum, while laser-power does not,
since in this particular case is marginal.

Let us consider the question of minimum power requirement as a function of payload mass
fraction . It is well known, that such a minimum exists for any thrust F, propulsion
requirement V and structural factor . However, at available exit velocities, has to be
kept far below optimum values, while on the other hand chemical power has no upper limit.

From equation (10) we derive:

dP gM n V 1 1
(20)
da a 1 1 a
n a n
a a
dP 1
(a) F (a) c(a) f (a)
da 2
(21)

dB
0! (22)
da

1
a a n 0! (23)
a

Table 1:
Data of laser-assisted chemical single-stage-to-orbit vehicles as a function of ratio of laser
power PL to chemical power PCh. There is in all cases a payload of 30.000 kg, structural factor
of 0.13 chemical exhaust velocity of 5000 m/s, take-off acceleration of 1.2 g and V of
10.000 m/s (tb = thrust time).

0 0.1 0.5 1 5 10 100


0.00533 0.01854 0.1131 0.3120 0.4172 0.4172 0.6895
Mo t 5622.9 1618.4 495.1 265.2 96.16 71.92 53.51
Ms t 731.0 210.4 59.7 34.5 12.50 9.35 5.66
Mp t 4861.9 1378.0 369.4 200.7 53.66 32.57 7.85
F 104N 6619 1905 540.5 312.2 113.2 84.7 51.2
c m/s 5000 5244 6114 7071 12247 16583 50249
13239 3633 882.6 441.5 92.43 51.05 10.19
m kg/s
tb s 367 379 418 455 580 638 770
P GW 165.5 49.95 16.55 11.04 6.932 7.019 12.865
PCh GW 165.5 45.41 11.03 5.52 1.155 0.638 0.1274
PL GW 0 4.54 5.51 5.52 5.777 6.381 12.7380

Table 2

Payload mass fraction cmin for minimum total power requirement as function of structural
factor .

0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.5 0.7 0.9 1.0
amin 0.367 0.365 0.357 0.347 0.334 0.320 0.304 0.287 0.229 0.143 0.049 0.0
9 0 7 4 7 3 4 2 8 5 3

Numerical solution of equation (23) results in the minimum values for shown in Table 2.

This minimum can, of course, easily be understood in terms of masses (see Fig. 6): power
requirement increases at low values of, due to raising Mo, at high values of it increases due to
decreasing propellant mass with accordingly increasing exhaust velocity requirements.
With a constant chemical contribution to power in laser-assisted chemical propulsion there,
too, should be minima for required laser power.

Differentiation of equation (17) gives:

2
dPL gM n 1 V CCh
(24)
da 2a a 1 V
n 2
a

2
1 V C Ch 1
( n )
a 1 V a
n
a

amin f ( V , CCh , ) (25)

An analytical expression in the form of equation (25) cannot be obtained by setting the factors
of equation (24) equal to zero. However, numerical solution of equation (24) or (17) gives the
desired answers (Fig. 5).
For a V = 10.000 m/s single-stage-to-orbit mission, the results are shown in Fig. 6. From
these, minimum required laser-power results with payload and acceleration as additional
parameters.

Equation (14) may be rewritten as follows:


2
gM n C Ch
PL ,min (c ) (26)
2a; min c

gM n F
PL ,min PL PL (27)
2a, min 2

2
C Ch V
PL' ,min c (28)
c 1
n
a min

2
C Ch 1
n
V a min

PL' being the equivalent laser-exhaust velocity. Plotted as a function of min with CCh and ß as
parameters (see Fig. 7), it delivers the required minimum laser power for any thurst level
(chosen according to equation (27) after selection of acceleration and payload along with
fixed min, the latter being determind by . V, CCh and ß.
As Fig. 7 shows, no min occurs where a line of constant CCh does not cross a particular line of
constant ß. Below a certain laser power simply decreases with decreasing . This is always
the case, when CCh alone is close to being sufficient for a particular mission, as in the example
of Table. 1.This is made more obvious by considering another sample case.

A single-stage-to-orbit ( V = 10.000 m/s) vehicle providing a chemical exhaust velocity of


CCh = 4500 m/s cannot be built with structural factors ß above 0.10837. In this case, it has
precisely zero payload. To design it with any other values of the sum + ß.requires additional
beamed power. The situation is depicted in Fig. 8.

PL is plotted over , with ß as parameter. All values of ß above 0.1 show minima at values
between 0.1 and 0.3. For very low values of , PL again decreases monotonously with (not
shown in Fig. 8). However, vehicles with very light construction (ß below 0.10837) require
smoothly raising amounts of laser power if increases beyond values where + ß equals
0.10837.
The PL lines of constant ß in Fig. 8 are very flat in the vicinity of this critical value. This again
reflects the fact found in the example of Table 1: in these regions a small increase in P L results
in a dramatic increase of possible payload massfractions .

In closing the present consideration of laser-assisted chemical propulsion, it should be


mentioned that, of course, all PL values considered here are effective values as they must
show up in the kinetic energy of the exhaust gases. Laser power to be transmitted from the
base power station PL,bas has to be larger according to transmission and conversion
efficiencies nT and nC, which normally will together be below 0.5:

PL nT nC PL,bas (29)

III. Transmission Considerations

Even in the case of short distance laser power transmission to ascending launch vehicles or to
spacecraft in low-earth orbits beam divergence, tracking and jitter may be well beyond
present-day technology. The situation is gravely worsened if transmission to geosynchronous
orbits, over lunar or interplanetary distances is considered. Due to beam divergence, power
desity is diluted by the square of the transmission distance. Since direct conversion to
propulsive power requires large total amounts of energy per second, which may not be
feasible with power-limited laser energy sources, indirect conversion by means of solar panels
might be of interest. This, as another advantage, offers the opportunity of making use of the
solar radiation as additional energy source, with which to compete any laser will have trouble
over larger distance. Electric power from the panels could then be converted to thrust with ion
or plasma engines.

Let us assume the existence of a laser-power base in low earth orbit, such that its distance
from the sun is essentially 1 AU (taken to be 1,496.1011 m).The base emits a conical laser
beam with divergence angle . The cross-sectional area of the beam at any distance from the
laser source DL will then be:

A DL2 tg 2 (30)
2

If the laser ha a power Eo at the source,its intensity will the diminish with DL:

Eo
IL (31)
2 2
D L tg
2

On the other hand, the solar intensity at 1 AU is ko. ko = 1,3146.103 W/m2.

The local solar intensity at any distance Ds from the sun is

2
AU
Is ko W / m2 (32)
Ds
The ratio of laser to solar intensity is therefore:

2
IL Ds 1 Eo
(33)
Is DL ko
AU 2 tg 2
2

The solar panel shall be a planarian surface of area A and radius r (not necessarily determined
by equation (30).The angle between this surface and the line of sight to the laser base be a, the
angle with the line of sight to the sun ß, so that the angle between two lines is the difference

= - (34)

The effective areas of the panel surface are then the projection upon a plane perpendicular to
the respective lines of sight:

AL r2 sin a (35)

As r2 sin (36)

sin sin (a ) sin a cos (37)

As r2 (sin a cos cos a sin ) . (38)

If laser beam divergence satisfies equation (30), the laser power received by the panel is

Eo
IL r2 sin a (39)
E
DL2 tg 2
2

while from the sun it receives

2
AU
Is ko r2 (sin a cos + (40)
Ds

+ cos a sin ).

Panel center, laser base and sun form a triangle with sides DL, AU and Ds, the angle formed
by DL and Ds. Therefore

AU 2 DL2 Ds2 2 DL Ds cos (41)

D L2 Ds2 AU
cos (42)
2 DL Ds
with the following abbreviations

2
AU
K1 ko r2 (43)
Ds

D L2 Ds2 AU 2
K2 (44)
2 DL Ds

DL2 Ds2 AU 2
K3 sin arc cos (45)
2 DL Ds

Eo
K4 r2 (46)
2 2
D L tg
2

the total power received by the panel becomes the following function of a:

I tot IL Is (47)

= ( K1 K 2 K 4 ) sin a K1 K 3 cos a.

Before studying the incluence of a, let us consider power and beam divergence requirements
as a function of distance.

For the simple case of the panel bearing spacecraft being in the direction opposite to the sun
(that is, the angle between the line of sight from base to sun and from base to spacecraft is
180 degress), the following results are obtained. It is further assumed that the panel has a size
given by local laser beam diameter and that the center line of the beam goes through the panel
center. is measured in fractions of one degree

a) Case of a 103 W laser

Fig. 9 plots I tot as a function of DL for this moderate size laser, with beam divergence as
parameter. Under the optimistic assumption of a pointing accuracy and beam divergence as
low as 10-5 (= 0.036 seconds!!) being feasible, this laser contributes substantially to solar
power only below 30.000 km distance.

Its influence at all distances beyond the earth-moon distance is negligible even if is
decreased by another order of magnitude.

Such a small laser is therefore of any interest only in the immediate vicinity of the earth, and
might assist operations of spacecraft within the shadow cone of the earth.
b) Case of a 106 W laser

Fig. 10: this laser is without any influence beyond lunar distances with = 10-5. However, it
might power electrical tugs to and from geosynchronous orbit even with larger beam
divergence.

It is a true but unrealistic statement that this laser, with 10-9 degrees beam divergence could
produce an illumination intensity equal to one solar constant ko as far as the distance of
Uranus. With equal pointing accuracy, the panel area required would have a diameter of 50 m.
An electrical engine with a specific impulse of 2000 s and a total panel- and engine-efficiency
of 20 % could produce a thrust of 20 N with a propellant consumption of 1 g/s.
C Case of a 109 W laser

Fig. 11: for all greater than 10-3degrees, the power intensity beyond lunar distance is
essentially ko, before it decreases below the level short of Mars due to decreasing solar
intensity. It would require 10-6 degrees to keep power level at 1 ko up to Mars. Jupiter distance
requires 10-7 degrees, Pluto 10-8.

Required panel diameter at Jupiter distance (and 10-7 degrees) is 1358 m. An electrical engine
could – under the same assumptions as mentioned above – produce a thrust of 2000 N at a
mass flow rate of 106 g/s.
As the above examples clearly demonstrate, the use of lasers over planetary distances is
completely unfeasible. If these cases were -limited, the situation becomes swiftly power-
limited when more realistic divergence angles are assumed.

With = 1 second (2,78.10-4 degrees), it takes 3.56 GW to produce the local solar intensity at
the distance of the moon, 64 200 GW for Mars in opposition and 516 000 for Pluto.

Since energy distribution across a laser beam as well as beam divergence are governed by
optical laws, including non-linear effects, none of the split-second divergence angles
considered above might ever become feasible. As Table 3 shows, it takes again multi-gigawatt
lasers to produce 1 ko only at the distance of the moon if an of 1.6710-2 degrees (1 minute) is
assumed:

Table 3
Range of lasers with power Eo to produce 1 ko intensity.

Eo(W) 106 107 108 109 12,2880.109


DL (km) 107 338 1070 3383 384.000

Therefore, once again, laser-powered propulsion will be restricted to short distances. At such
distances, if indirect conversion is used rather than directly laser-powered engines, the
simultaneous use of solar light is most rewarding. (For direct conversion propulsion, the
additional use of chemical power at appropriate V/CCh-missions leads to moderate laser
power requirements. See previous chapter).

Turning back to equation (47) we find:

dE tot
( K1 K 2 K 4 ) cos a K1 K 3 sin a (48)
da

K1 K 2 K 4
amax = arc tg (49)
K1 K 3

amax being the angle between panel surface and line of sight to the laser base for obtaining
maximum combined solar and laser energy input.

Although equation (49) is valid for any value of the constants, the optimization of at
moderate laser powers is worthwhile only at rather short distances.

Fig. 12 shows the result for a 1000 W laser with = 10-6 (which again is unrealistic, but the
result is quite typical): for all distances above 40.000 km, the panel has always to be turned
towards the sun. It is a question of payload gain or thrust-to-weight gain, whether the required
amount of attitude control pays off.
IV. Conclusion

Laser propulsion is worthwhile for high V missions only. With limited laser power, laser
assisted chemical propulsion can lead to very high payload gains over purely chemical
propulsion. This would require the use of chemical propellant combinations rather than inert
materials. Long-range laser propulsion requires very low beam divergence angles. Laser
propulsion will therefore have to be confined to near-earth missions. Two-way tugs and
single-stage-to-orbit shuttles are attractive candidates.

However, these and other missions are up to several orders of magnitude beyond present-day
technology in one or several of the following areas.

Laser power output


Mirror materials for very hig intensity energy fluxes
Solar panels for very high intensity energy fluxes
Pointing accuracy
Jitter reduction
Beam divergence angle.

However, it is hoped that the arguments presented above contribute to the conviction shared
by many scientists not afraid of looking into a somewhat more distant future, that laser
propulsion is a rewarding field for further studies.

Anda mungkin juga menyukai