Fact-checked by Grok 2 weeks ago

Relativistic rocket

A relativistic rocket is a theoretical system for a that achieves velocities approaching the , where must be applied to accurately describe its dynamics, momentum conservation, and energy expenditure, contrasting with classical Newtonian rocket equations that fail at such speeds. This concept addresses the challenges of by modeling how a expels reaction mass or photons at relativistic speeds, leading to significant effects like , , and increased effective mass. The foundational relativistic rocket equation, derived by Jakob Ackeret in 1946, expresses the attainable v in terms of the exhaust velocity v_e and the m_0 / m_f (initial to final mass) as \frac{1 + v/c}{1 - v/c} = \left( \frac{m_0}{m_f} \right)^{2 v_e / c}, incorporating Lorentz factors to account for the relativistic addition of velocities and the transformation of momentum. This equation reveals that achieving even modest fractions of the requires exponentially larger mass ratios compared to non-relativistic cases, highlighting the immense fuel demands for high-speed . Subsequent refinements, such as those by G. R. Purcell in 1966 and Ulrich Walter in 2006, extended the model to include photon exhaust and proper speed metrics for better parameterization of relativistic trajectories. Key implications of relativistic rocketry include the asymmetry between acceleration and deceleration phases due to the invariance of the speed of light, as well as the potential for hyperbolic motion under constant , where the rocket experiences a steady in its instantaneous . These factors make relativistic rockets central to theoretical designs for or laser-driven systems aimed at nearby stars, though practical realization remains constrained by energy densities far beyond current . Overall, the framework underscores the fundamental limits imposed by on , emphasizing the need for advanced fuels like to approach viable interstellar velocities.

Theoretical Foundations

Core Principles of Special Relativity in Propulsion

In the context of propulsion systems approaching significant fractions of the speed of light, the principles of special relativity fundamentally alter the classical notions of space, time, and motion, requiring a framework that accounts for the invariance of the speed of light and the relativity of simultaneity. Lorentz transformations provide the mathematical foundation for describing how coordinates and times of events differ between inertial frames, particularly relevant for a rocket transitioning between rest and high velocities relative to an Earth-based observer. For a rocket moving along the x-axis with velocity v relative to an inertial frame S, the Lorentz transformations relating coordinates (ct, x, y, z) in S to (ct', x', y', z') in the rocket's instantaneous rest frame S' are given by: \begin{align} ct' &= \gamma (ct - \beta x/c), \\ x' &= \gamma (x - vt), \\ y' &= y, \\ z' &= z, \end{align} where \beta = v/c, \gamma = 1/\sqrt{1 - \beta^2}, and c is the . These transformations imply key effects in rocket motion: , where the proper time \tau elapsed on the rocket is \tau = t / \gamma shorter than t observed from , slowing aging for the crew; , contracting the rocket's length in the direction of motion to L = L_0 / \gamma as seen from , where L_0 is the ; and relativistic velocity addition, preventing velocities from simply summing classically. If the rocket ejects exhaust at speed u' relative to itself, the exhaust speed u observed from is u = (v + u') / (1 + vu'/c^2), ensuring no speed exceeds c. A critical distinction in relativistic propulsion arises between proper acceleration, the acceleration measured in the instantaneous of the (felt by the crew as constant, often denoted \alpha), and coordinate acceleration (the rate of change of velocity as observed from a distant inertial , denoted a = dv/dt). remains invariant across frames and is related to coordinate acceleration by \alpha = \gamma^3 a, meaning that as \gamma increases with speed, the same felt onboard corresponds to diminishing coordinate acceleration from , approaching zero asymptotically near c. This —termed hyperbolic motion—results from constant , where the 's worldline in follows a , bounding the observable universe for the crew due to an forming behind the spacecraft. The mass-energy equivalence principle, E = mc^2, underscores the immense requirements for relativistic speeds, as must be converted into with near-perfect to achieve significant \gamma. In propulsion, expelling mass dm at high exhaust velocity imparts , but the total energy released from conversion equals the rest energy mc^2 of the annihilated or reacted , limiting practical designs to exotic mechanisms like matter-antimatter reactions where up to 100% of converts to directed . Without such , the mass fraction needed grows exponentially, rendering infeasible under classical limits but theoretically viable within relativistic constraints. Early theoretical foundations for relativistic rocket motion trace to Max Born's 1909 analysis of under constant , introducing the concept of hyperbolic motion to maintain structural integrity in accelerated frames. This work was refined in the mid-20th century, notably by Wolfgang Rindler in the 1950s and 1960s, who developed coordinates (now ) describing uniformly accelerating observers and their implications for propulsion, emphasizing the role of in avoiding paradoxes like infinite energy demands.

Distinction from Classical Rocketry

The classical , \Delta v = v_e \ln(m_0 / m_f), where \Delta v is the change in , v_e is the exhaust , m_0 is the , and m_f is the final , relies on assumptions rooted in Galilean and Newtonian . This formulation treats as simply p = m v with constant inertial and neglects any speed-dependent effects on energy or , making it valid only for velocities much less than the (v \ll c). At low speeds, such as those achieved by chemical rockets (typically v < 0.01c), it accurately predicts fuel requirements for missions like orbital insertion or interplanetary travel. However, the classical equation fails dramatically at velocities exceeding 0.1c, where relativistic effects become significant. For instance, it predicts finite energy needs to approach c, but in reality, accelerating any massive object to the speed of light requires infinite energy due to the divergence of relativistic effects. This breakdown arises because the equation ignores the increase in the rocket's inertial mass with velocity, leading to underestimated propellant demands and overestimated achievable speeds. Energy efficiency also plummets in the relativistic regime, as a larger fraction of the input energy goes into overcoming the growing inertia rather than pure kinetic gain, resulting in rapidly diminishing returns for additional fuel expenditure. A quantitative comparison illustrates this divergence starkly. The Lorentz factor, \gamma = 1 / \sqrt{1 - v^2/c^2}, underlies the discrepancy, growing from near 1 at low speeds to infinity as v \to c. At 0.99c, where \gamma \approx 7.09, the relativistic mass ratio required for a given exhaust velocity is orders of magnitude higher than the classical prediction—for example, to reach 0.99c with v_e = 0.1c, a classical calculation suggests a mass ratio of approximately 20,000, while the relativistic case requires about $3 \times 10^{11}, rendering such designs impractical with foreseeable technology due to the exponential fuel scaling. This contrast highlights why classical models cannot guide designs for interstellar propulsion at near-light speeds.

Relativistic Rocket Equation

Derivation and Formulation

The relativistic rocket equation is derived from the conservation of four-momentum in special relativity, applied to a rocket that ejects exhaust mass at a constant relative velocity u (the effective exhaust velocity v_e = u) backward in its instantaneous rest frame. Consider the rocket with current rest mass m moving at velocity v in an inertial reference frame. In a small time interval, it ejects a small rest mass dm > 0, reducing its rest mass by dm. The velocity of the ejected mass in the inertial frame is obtained via relativistic velocity addition: v_{\text{ex}} = \frac{v - u}{1 - \frac{v u}{c^2}}, where c is the . The Lorentz factor for the exhaust is \gamma_{\text{ex}} = \gamma_v \gamma_u \left(1 - \beta_v \beta_u\right), with \gamma_v = \left(1 - \frac{v^2}{c^2}\right)^{-1/2}, \gamma_u = \left(1 - \frac{u^2}{c^2}\right)^{-1/2}, \beta_v = v/c, and \beta_u = u/c. Conservation of energy in the inertial frame gives d(\gamma_v m c^2) = - \gamma_{\text{ex}} dm \, c^2, or equivalently, d(\gamma_v m) = - \gamma_{\text{ex}} \, dm. (along the direction of motion) gives d(\gamma_v m v) = - \gamma_{\text{ex}} v_{\text{ex}} \, dm. Substituting the expression for d(\gamma_v m) into the momentum equation yields \gamma_v m \, dv + v \, d(\gamma_v m) = v_{\text{ex}} \, d(\gamma_v m), which simplifies to \gamma_v m \, dv = (v_{\text{ex}} - v) \, d(\gamma_v m). The term v_{\text{ex}} - v = - \frac{u}{\gamma_v^2 (1 - \beta_v \beta_u)}. The \gamma_v factors from the Lorentz transformations and laws cancel in the rearrangement, leading to a separable . To solve, introduce the rapidity \phi = \artanh(v/c), so v = c \tanh \phi and dv = c \sech^2 \phi \, d\phi = c (1 - v^2/c^2) d\phi = (c^2 / \gamma_v^2 v) dv wait, actually, since \gamma_v = \cosh \phi and the differential form becomes d\phi = \frac{u}{c} \frac{dm}{m} (noting dm < 0 for the rocket mass change). Integrating from initial mass m_0 (at v=0, \phi=0) to final mass m_f gives \phi = \frac{u}{c} \ln \left( \frac{m_0}{m_f} \right). Thus, the final velocity is v_f = c \tanh \left[ \frac{u}{c} \ln \left( \frac{m_0}{m_f} \right) \right]. This form arises because rapidity adds linearly under successive velocity boosts, analogous to velocity addition in classical mechanics, but bounded by the hyperbolic tangent to ensure v_f < c. The original derivation of this equation was provided by Ackeret in 1946, with a transparent presentation using rapidity by Forward in 1995. The "relativistic mass ratio" refers to the ratio m_0 / m_f, which plays the same role as in the classical Tsiolkovsky equation v_f = u \ln (m_0 / m_f). However, unlike the classical logarithmic scaling (where v_f is unbounded), the relativistic version uses the tanh function, asymptotically approaching c as the mass ratio increases, reflecting the Lorentz invariance and speed limit of special relativity. This difference highlights how relativistic effects modify propulsion efficiency at high speeds, requiring exponentially larger mass ratios to achieve velocities near c. An alternative formulation uses directly: \phi = \artanh(v/c) = (u/c) \ln (m_0 / m_f), emphasizing the additive nature of rapidity boosts from each exhaust ejection. For the case of constant \alpha (the acceleration measured in the instantaneous ), the equation integrates over \tau: v = c \tanh(\alpha \tau / c). Here, \alpha = c^2 / R, where R is the of the rocket's worldline in , describing hyperbolic motion. This form assumes adjusted to maintain constant \alpha as mass decreases, distinct from constant exhaust velocity.

Physical Interpretations and Limits

The relativistic rocket equation implies that the ship's velocity v approaches the c asymptotically as the (initial mass to final mass) increases, but never reaches or exceeds c, due to the infinite energy required at that limit. This behavior arises from the \gamma = 1 / \sqrt{1 - v^2/c^2}, which diverges as v \to c, reflecting the fundamental constraint of on massive objects. In practical terms, even with an arbitrarily large , the ultimate velocity remains just below c, bounded by the rocket's structural limits at the Planck scale for subatomic components. The underscores the immense demands of relativistic . The total of the is given by E = \gamma m c^2, where m is the instantaneous rest mass and \gamma accounts for both the rest m c^2 and the (\gamma - 1) m c^2. Consequently, the must supply not only the to accelerate the but also compensate for the rest of the expended , which is ejected relativistically and carries away significant . This dual requirement amplifies the effective needed compared to classical rocketry, where costs are Newtonian. A concrete illustration of these limits is a journey to the , approximately 2.5 million light-years distant, under constant \alpha = 1g \approx 9.81 \, \mathrm{m/s^2}. From the ship's perspective, the elapsed is about 28 years, allowing the crew to experience the trip within a human lifetime, while from Earth's inertial frame, the is roughly 2.5 million years, as the ship covers the distance at speeds nearing c. Relativistic effects also manifest in an "" analogous to that in for the accelerating frame. From the Earth's viewpoint, the ship appears to slow asymptotically toward c, with signals from beyond a certain rearward distance unable to catch up due to the horizon at -c^2 / \alpha behind the ship; objects crossing this horizon redshift and fade indefinitely. However, the crew experiences a finite with no such slowing, highlighting the frame-dependent nature of the journey's duration and observability.

Advanced Propulsion Concepts

Matter-Antimatter Annihilation Drives

Matter-antimatter drives harness the complete conversion of into , achieving 100% in transforming the rest of annihilating particles and antiparticles into and , far surpassing other methods. In electron-positron , an (e⁻) and positron (e⁺) collide to produce two gamma-ray photons:
e^- + e^+ \to 2\gamma
releasing a total of 1.02 MeV (0.511 MeV per photon). For relativistic rocket applications, proton-antiproton (p¯p) is more practical, yielding approximately 1.88 GeV per through the of pions and other particles, with about 40% of the carried by charged pions that can be directed for . This process delivers an of $9 \times 10^{16} J/kg when equal masses of and are annihilated, equivalent to the full E = mc^2 conversion.
The drive configuration involves injecting into a chamber where occurs, producing high-energy products such as gamma rays, neutral pions (decaying into gamma rays), and charged pions (π⁺ and π⁻). Charged pions, moving at near-light speeds, are collimated and directed rearward using magnetic nozzles to generate , while gamma rays pose challenges due to their isotropic and require reflective or absorptive materials for partial utilization. This setup minimizes the need for additional by leveraging the products themselves as the exhaust, enabling exhaust velocities approaching the (v_e \approx c). In ideal conditions, the (I_{sp}) for such drives reaches approximately $3 \times 10^7 seconds, calculated as I_{sp} \approx v_e / g_0, where g_0 \approx 9.8 m/s² is Earth's , allowing relativistic velocities with minimal penalties under the relativistic rocket . This high I_{sp} stems from the near-c exhaust speed, making annihilation drives suitable for missions requiring speeds of 0.1c to 0.5c. A seminal concept is the beam-core antimatter rocket proposed by in the 1980s, where streams of s are injected into a proton-rich to initiate , producing charged pions that are magnetically focused into a for direct generation. In this design, the engine achieves an I_{sp} of about 28 million seconds with levels on the order of tens of newtons, using antiproton flow rates of 100 micrograms per second, optimizing for both efficiency and practicality in relativistic propulsion.

Pion-Based Rocket Designs

Pion-based rocket designs leverage the charged s generated from antiproton-proton as the primary exhaust for in relativistic . In a typical antiproton-proton at rest, the reaction yields an average of about 3.2 charged s (a mix of π⁺ and π⁻) alongside neutral pions, with each charged possessing a total energy of approximately 395 MeV, corresponding to kinetic energies of roughly 250 MeV after accounting for the 's rest mass of 140 MeV. These pions carry a significant fraction—around 40%—of the total energy (1.876 GeV), making them a high-velocity reaction product suitable for generation without additional . The core of these designs involves onboard annihilation of antiprotons with protons, followed by immediate channeling of the charged pions using to produce directed . Seminal concepts from in the 1980s proposed employing a or mirroring to separate the charged pions from neutral particles and gamma rays, directing them aftward to exploit their relativistic speeds (near 0.94c) for exhaust velocities exceeding 0.9c. This approach achieves efficiencies of up to 70% by minimizing energy losses to isotropic emissions, far surpassing classical rocket limits while enabling specific impulses on the order of 10^7 seconds. Mass ratio analyses have demonstrated feasibility for cruise velocities up to 0.9c, with payload fractions remaining viable despite relativistic effects on the rocket equation. A key challenge in pion-based systems stems from the short mean lifetime of charged pions, approximately 26 ns at rest (dilated relativistically during ), necessitating annihilation and acceleration occur within a compact reaction chamber to prevent into muons and neutrinos before ejection. This requires precise magnetic confinement to rapidly guide the pions—produced isotropically—into a , with onboard antiproton storage and proton target integration adding complexity but enabling continuous operation. While these designs build on the broader framework of matter-antimatter for energy release, their focus on pion directionality optimizes for missions.

Engineering Challenges

Energy and Efficiency Constraints

Achieving relativistic speeds imposes severe energy demands on systems, primarily due to the increase in required as velocities approach the . For a maintaining constant of 1 g to reach 0.99c, the requirements escalate dramatically; estimates for a multi-ton indicate peak outputs on the order of 10^{12} to 10^{14} W, far exceeding current capabilities of around 10^9 W and comparable to or surpassing global human production, which averaged about 20 TW as of 2024. These demands arise from the need to continuously supply against increasing relativistic , with in the laboratory frame scaling as P \approx m \alpha v (approaching m \alpha c at high velocities), where \alpha is and m is rest , leading to terawatt-level engines for practical mission masses of 1000–10,000 tons. Antimatter, prized for its unparalleled of nearly $9 \times 10^{16} J/kg upon with , remains the most viable candidate, yet its and present insurmountable near-term barriers. Theoretical costs for antiprotons exceed $6 \times 10^{15} per gram based on energy input and low (around 10^{-9}), translating to over $6 \times 10^{18} per kg, rendering even milligram quantities prohibitively expensive. relies on Penning traps, which use combined electric and to confine charged plasmas at densities up to 10^{11} particles per trap, but scaling to kilogram levels would require vast arrays of such devices, consuming gigawatts continuously and risking upon contact with container walls. Updated assessments from and in the 2020s confirm that annual antimatter output hovers at 1–10 ng, primarily antiprotons from the Antiproton Decelerator, insufficient even for micro-thrust validation in space propulsion tests and highlighting a production bottleneck that would take billions of years at current rates to yield 1 g. This scarcity underscores the scalability constraints, as diverting full accelerator capacity to propulsion would still yield only nanograms annually, far below the kilograms needed for meaningful relativistic missions. Efficiency further compounds these challenges, with annihilation drives suffering significant losses from nozzle inefficiencies and radiative cooling. In beam-core configurations, where charged pions from proton-antiproton provide , isotropic emission and magnetic nozzle divergence limit the effective exhaust velocity (v_e) to 0.3–0.5c, despite theoretical pion speeds near 0.94c, resulting in overall efficiencies of 30–50% due to uncollimated gamma rays and escape. These losses necessitate oversized fuel reserves, exacerbating the already daunting and mass penalties inherent to relativistic flight.

Material and Stability Issues

At relativistic speeds, experience significant structural stresses due to differential across their length. To maintain structural integrity under constant , different sections of an extended vehicle must accelerate at slightly varying rates, leading to tidal-like forces that induce tensile stress in the frame. This effect, illustrated in analyses of , requires precise coordination to prevent material failure, as uncoordinated acceleration would cause elongation and potential rupture of connecting elements. External relativistic stresses further compound these challenges through interactions with . Blueshifted cosmic rays, primarily protons and alpha particles, increase in energy and flux as the γ rises, with the dose rate scaling approximately as γ³ due to beaming, , and increased particle density in the forward direction. For instance, at γ ≈ 10, the flux can exceed 10¹³ particles m⁻² s⁻¹, delivering lethal doses without protection. Advanced shielding is essential, such as multi-meter-thick or layers, or hybrid combined with material barriers like iron, to deflect and absorb these high-energy particles and mitigate hazards. Maintaining stability during prolonged constant proper acceleration demands robust vibration control to avoid structural fatigue. High-frequency vibrations from engine thrust or relativistic effects can amplify under the equivalent gravitational field (e.g., proper acceleration), leading to cyclic loading that weakens materials over time. NASA studies on space vehicle dynamics emphasize predictive modeling of vibration modes to design damping systems, ensuring resonance is avoided in extended missions. Recent advancements in materials address these durability issues, particularly for withstanding high-energy particle impacts. In the 2020s, NASA's Superlightweight Aerospace Composites project has developed (CNT)-reinforced composites that offer superior strength-to-weight ratios, enabling lighter structures capable of deep space environments. These materials, integrated into shielding composites, provide effective protection against high-energy particles, including those analogous to pions in propulsion systems, by dissipating impact energy without significant degradation. Thermal management poses additional stability risks from high-energy processes like matter-antimatter annihilation, which releases immense (up to $9 \times 10^{13} J/g) primarily as gamma rays and charged particles. Without adequate dissipation, this can cause or of components, compromising structural integrity. Solutions include coolants, such as sodium-potassium alloys, circulated through microchannels to absorb and transfer at temperatures exceeding 500 K, as explored in propulsion designs. Complementary via large deployable radiators rejects to space, with concepts emphasizing high-emissivity surfaces to handle the loads in antimatter rockets.

References

  1. [1]
    [PDF] arXiv:1504.07205v1 [physics.space-ph] 27 Apr 2015
    Apr 27, 2015 · The relativistic rocket equation was first derived by Ackeret ([2]) in 1946, and subsequently expounded upon by Krause ([9]), Bade ([6]), and ...
  2. [2]
    [PDF] AS T RO.RE LATIVI TY - NASA Technical Reports Server
    Generalized relativistic conservation laws of momentum and total energy. (mass) are derived. An application of relativistic dynamics to rocket propulsion gives ...
  3. [3]
  4. [4]
    Relativistic rocket and space flight - ADS - Astrophysics Data System
    The rocket equations are even identical. In addition, we generalize the expressions for the rocket exhaust speed to systems with mass and/or photon propulsion.<|control11|><|separator|>
  5. [5]
    [PDF] Pros and cons of relativistic interstellar flight - arXiv
    Owing to its highest energy density, antimatter in its condensed state (liquid or solid) is the most relevant fuel for powering a multi-ton relativistic rocket ...
  6. [6]
    [PDF] Antimatter Propulsion - NASA Technical Reports Server
    Antimatter has high energy density (~9x10^10 MJ/kg), 10 billion times more than chemical combustion, and opposite charge/magnetic moment compared to matter.
  7. [7]
    [PDF] The Lorentz Transformation - UCSB Physics
    To derive the Lorentz Transformations, we will again consider two inertial observers, moving with respect to each other at a velocity v. This is illustrated in ...
  8. [8]
    [PDF] The theory of the rigid electron in the kinematics of the principle of ...
    It is remarkable that no matter how large its acceleration might be, an electron in hyperbolic motion will emit no actual radiation, but its field will move ...
  9. [9]
    The Relativistic Rocket
    At the trip's end the fuel has all been converted to light with momentum of magnitude EL/c, but in the opposite direction to the rocket. So the final ...
  10. [10]
  11. [11]
    Nearlight Starships - Atomic Rockets
    Aug 24, 2022 · At 30% of c, the difference between relativistic mass and rest mass is only about 5%, while at 50% it is 15%, (at 0.75c the difference is over ...
  12. [12]
    A transparent derivation of the relativisitic rocket equation | Joint Propulsion Conferences
    Insufficient relevant content. The provided URL (https://arc.aiaa.org/doi/pdf/10.2514/6.1995-3060) contains only a title, metadata, and navigation elements, with no accessible text or derivation of the relativistic rocket equation. No step-by-step derivation, key equations, or concepts like four-momentum, rapidity, or relativistic mass ratio are present in the visible content.
  13. [13]
    [PDF] Dynamics and Relativity: Example Sheet 4 - DAMTP
    Write down the Lorentz transformation law for the energy and relativistic 3- momentum of a particle. E′ = γE(1 − β cosθ) and cosθ′ = cosθ − β 1 − β cos θ ,
  14. [14]
    The Relativistic Rocket
    ### Summary of Relativistic Rocket Content
  15. [15]
    The Ultimate Limits of the Relativistic Rocket Equation. The Planck ...
    Jul 26, 2018 · We are basically combining the relativistic rocket equation with Haug's new insight on the maximum velocity for anything with rest mass. An ...
  16. [16]
    The Rindler Horizon - Greg Egan
    Sep 20, 2006 · Because we're dealing with relativistic velocities, we need to be precise and say that the spaceship undergoes constant proper acceleration: ...
  17. [17]
    Future of antimatter production, storage, control, and annihilation ...
    Utilizing antimatter annihilation, this propulsion method boasts an unmatched energy density of 9 × 10¹⁶ J/kg, released with 100% efficiency when antimatter ...
  18. [18]
    [PDF] Antiproton Annihilation Propulsion - DTIC
    Antiproton annihilation propulsion uses antimatter to heat reaction fluid, which is then exhausted to produce high thrust. It is feasible but expensive.
  19. [19]
    Antiproton annihilation propulsion - AIAA ARC
    On the average there are 3.2 charged pions and 1.6 neutral pions.2 The neutral pions have a lifetime of only 90 at- toseconds and almost immediately convert ...Missing: proton | Show results with:proton
  20. [20]
    How is the energy distributed in a proton-antiprotion annihilation?
    Aug 13, 2013 · Firstly, the neutral pion gets 390 MeV total energy, of which 135 MeV is the rest mass energy and 255 MeV is its kinetic energy.
  21. [21]
    Deep Space Propulsion
    Ideally any rocket would have a lower mass ratio and so have less propellant, ... American aerospace engineers Robert Zubrin and Dana Andrews [2, 3] a sail.
  22. [22]
    Charged-Pion Lifetime and a Limit on a Fundamental Length
    The lifetime of the charged pion has been determined in flight. The agreement of our result, 26.02 ± 0.04 nsec, with experiments of comparable precision.Missing: rocket design
  23. [23]
    [PDF] Antiproton Powered Propulsion With Magnetically Confined Plasma ...
    Antiproton propulsion uses matter-antimatter annihilation, where antiproton beams annihilate with hydrogen gas, transferring energy to propellant for thrust.<|control11|><|separator|>
  24. [24]
    Energy Production and Consumption - Our World in Data
    averaging around 1% to 2% per year. Primary energy consumption. Total energy ...
  25. [25]
    [PDF] Antimatter Requirements and Energy Costs for Near-term ...
    Substituting these values and an assumed kg,.id of. $0.10 per kW-hr into Eq. (4) yields an energy cost of $6,410 trillion per gram of antiprotons. Obviously ...Missing: 2020s | Show results with:2020s
  26. [26]
    The Antiproton Decelerator - CERN
    The Antiproton Decelerator (AD) is a unique machine that produces low-energy antiprotons for studies of antimatter, and “creates” antiatoms.Missing: 2020s ng/
  27. [27]
    Making antimatter | Angels & Demons - The science behind the story
    Even if CERN used its accelerators only for making antimatter, it could produce no more than about 1 billionth of a gram per year.Missing: 2020s | Show results with:2020s
  28. [28]
  29. [29]
  30. [30]
    [PDF] STRUCTURAL VIBRATIONPREDICTION
    This report focuses on predicting structural vibration in space vehicles, determining internal loads and stresses from induced or natural environments.
  31. [31]
    Superlightweight Aerospace Composites - NASA
    Jan 21, 2021 · The SAC project advances the maturation of Carbon Nanotube (CNT) reinforced composites to enable their application in aerospace structures.
  32. [32]
    Shielding of Cosmic Radiation by Fibrous Materials - MDPI
    Oct 15, 2021 · We give an overview of fiber-based shielding materials, mostly applied in the form of composites, and explain why these materials can help shielding spaceships ...