Fact-checked by Grok 2 weeks ago

Expendable launch system

An expendable launch system, also known as an expendable launch vehicle (ELV), is a type of designed for single-use operation, in which its propulsive stages are not recovered or reused after delivering a to or beyond. These systems discard components post-separation, often resulting in stages falling into the ocean or disintegrating during atmospheric re-entry. Historically, expendable launch vehicles have dominated space access, enabling the deployment of satellites, interplanetary probes, and crewed mission elements through programs like NASA's and Titan/Centaur boosters, which supported operations for nearly three decades. Their design simplicity—lacking recovery hardware—facilitates potentially lower production costs and allows full utilization for , yielding higher payload fractions compared to reusable counterparts that reserve resources for . The U.S. Evolved Expendable Launch Vehicle (EELV) program exemplifies ongoing reliance on such systems for assured, reliable access to space via vehicles like and . While reusability promises cost reductions over multiple flights, expendables persist for missions prioritizing maximum performance and proven reliability, though they generate significant debris and per-launch expenses exceeding tens of millions of dollars.

Definition and Fundamentals

Definition and Classification

An expendable launch system, also known as an expendable launch vehicle (ELV), consists of a or vehicle engineered to transport a into during a single mission, after which its stages and major components are neither recovered nor reused. The propulsive elements, including boosters and upper stages, are designed to be expended—operated once and then discarded, often disintegrating upon atmospheric reentry or remaining as . This single-use architecture prioritizes mission reliability and efficiency over cost from hardware , a approach dominant in space access since the due to the high structural stresses of launch and the complexities of safe . ELVs typically employ multi-stage designs with chemical , where each stage ignites sequentially to overcome and achieve orbital , jettisoning empty stages to reduce mass. Expendable launch vehicles are classified primarily by payload capacity to a reference , such as a 100 (nm) circular orbit, which determines their suitability for small satellites, crewed missions, or heavy interplanetary probes. U.S. federal regulations under 14 CFR § 420.19 delineate weight classes based on maximum in pounds (lbs) for launches at 28° and 90° orbital inclinations, accounting for launch latitude effects on performance.
Class28° Inclination Payload (lbs)90° Inclination Payload (lbs)
Small≤ 4,400≤ 3,300
Medium> 4,400 to ≤ 11,100> 3,300 to ≤ 8,400
Medium-Large> 11,100 to ≤ 18,500> 8,400 to ≤ 15,000
Large> 18,500> 15,000
These categories align with industry conventions of small-lift (under ~2,000 kg to ), medium-lift (~2,000–20,000 kg), and heavy-lift (over ~20,000 kg) vehicles, though exact thresholds vary by reference and efficiency. Additional subclassifications may consider stage count—typically two to four—or configuration variants, such as adding solid rocket boosters to liquid-fueled cores for enhanced thrust.

Core Operational Principles

Expendable launch systems (ELVs) operate by sequentially igniting and discarding multiple rocket stages to achieve the high velocities required for orbital insertion or beyond, typically around 7.8 km/s for . Each stage consists of engines, propellants, and structural elements tailored for a specific of ascent, with chemical —either solid or liquid—providing the . Upon depletion in a stage, separation occurs via pyrotechnic bolts, pneumatic systems, or springs, shedding inert mass to enhance efficiency for the upper vehicle assembly. This staged discard is essential to overcome the exponential propellant demands dictated by the , where delta-v capability is proportional to exhaust velocity times the natural log of the initial-to-final ; staging effectively resets the for each segment. The launch sequence begins with pre-ignition checks, arming of ordnance, and engine start on the pad, ensuring thrust-to-weight ratio exceeds unity for liftoff. Initial ascent is near-vertical to clear the launch tower and tower, followed by a programmed pitch-over into a that leverages aerodynamic lift and Earth's curvature to build horizontal velocity while conserving energy. Critical events include maximum (Max-Q), typically 1-2 minutes post-liftoff, where vehicle speed and atmospheric density combine to impose peak aerodynamic loads, necessitating throttled engines or structural reinforcements. Subsequent events propel the through exo-atmospheric phases, with the —protecting against ascent heating and drag—jettisoned once above ~100 km altitude. Upper stages execute precise burns under inertial guidance, often augmented by star trackers or GPS, to circularize orbits or inject into geostationary transfer orbits (GTO). Payload separation follows via clamps or yo-yo de-spin mechanisms for spin-stabilized upper stages, deploying satellites or probes into their operational trajectories. Unlike reusable systems, ELVs incorporate no recovery parachutes, heat shields, or retro-propulsion, allocating mass savings to greater propellant loads or larger payloads, which historically enables ELVs to deliver up to 18,000 kg to LEO in configurations like the Titan IV. Post-mission, expended stages follow ballistic trajectories, with boosters impacting remote ocean areas and upper stages either deorbiting or remaining in heliocentric orbits, contributing to space debris concerns.

Historical Development

Origins in Military Ballistics

The development of expendable launch systems originated from military ballistic missile programs, which prioritized reliable vertical ascent and high-velocity propulsion for delivering warheads over long ranges. The German V-2 (Vergeltungswaffe 2), engineered under Wernher von Braun, marked the first operational large-scale liquid-propellant rocket, achieving suborbital flight with a maximum altitude of approximately 189 kilometers on test flights. Initial static tests began in 1941, followed by the first full launch attempt on June 13, 1942, which failed due to propellant feed issues; the inaugural successful flight occurred on October 3, 1942, reaching 84.5 kilometers. By late 1944, over 3,000 V-2s had been combat-launched against Allied cities, demonstrating vertical takeoff, inertial guidance, and separation of engine from payload, though with a failure rate exceeding 20 percent due to production haste and quality control lapses under wartime constraints. These missiles, inherently expendable as single-use weapons, laid the engineering groundwork for space launch vehicles by proving scalable liquid oxygen-ethanol propulsion and aerodynamic stability during ascent. Following , Allied powers captured V-2 components, blueprints, and personnel, repurposing the technology for both military and scientific rocketry. In the United States, relocated von Braun's team to develop the missile, an (IRBM) with a range of 200-250 miles, first statically tested in 1952 and launched operationally on August 20, 1953, from . The , retaining the V-2's single-stage expendable architecture but with improved guidance and a cluster of control thrusters, directly supported early space efforts, including suborbital tests and the that carried on the first American crewed suborbital flight on May 5, 1961. Meanwhile, the reverse-engineered captured V-2s into the R-1 missile, with its first successful launch on September 17, 1948, evolving toward the ICBM, a clustered-engine design first tested on May 15, 1957. These military programs inherently favored expendable designs for cost-effective and mission-specific optimization, contrasting with later reusable concepts, as ballistic trajectories demanded high-thrust, one-time-burn engines without recovery mechanisms. The R-7's adaptation for orbital insertion—without major structural changes—enabled the launch of on October 4, 1957, the first artificial satellite, underscoring how ICBM reliability (achieved through redundant strap-on boosters) directly translated to launch vehicle performance. U.S. efforts similarly progressed from V-2-derived sounding rockets, with 86 captured units launched between 1945 and 1952 at White Sands, to multi-stage hybrids like the Bumper (V-2 with upper stage) first fired on February 24, 1949, reaching 400 kilometers altitude. This ballistic heritage imposed limitations, such as boil-off and non-recoverable staging, but established core principles of payload separation and velocity buildup essential for orbital insertion.

Space Race and Early Orbital Launches

The initiated the orbital phase of the on October 4, 1957, with the launch of , the first artificial satellite, aboard an rocket variant from the . The , originally developed as an (ICBM) with a clustered design of four strap-on boosters around a central core, achieved for the 83.6 kg payload after a two-stage ascent, marking the debut of an expendable launch system capable of orbital insertion. This single-use vehicle discarded its stages sequentially, with no recovery mechanisms, prioritizing payload delivery over reusability due to the era's propulsion and materials limitations. In response, the accelerated its efforts following two failed Vanguard attempts in late 1957, successfully orbiting on January 31, 1958, via a rocket from . The , a four-stage expendable vehicle derived from the Jupiter-C sounding rocket and ballistic missile lineage, lofted the 13.97 kg satellite into an elliptical with a perigee of 358 km and apogee of 2,531 km, confirming the Van Allen radiation belts through its instrumentation. Like the R-7, it employed irreversible stage separation and atmospheric reentry for discarded components, reflecting the causal priority of achieving orbit amid geopolitical urgency rather than cost recovery. These pioneering launches spurred rapid iterations, with the R-7 family enabling subsequent Soviet milestones, including Yuri Gagarin's manned orbital flight on April 12, 1961, while the U.S. transitioned to Atlas and expendables for Mercury and early scientific missions. Both nations' systems, rooted in wartime ballistic heritage yet adapted for sustained orbital access, demonstrated high reliability through serial production— the R-7 achieving over 90% success in early variants— but underscored expendability's trade-offs in resource intensity, as each mission consumed millions in materials without salvage. This era established expendable architectures as the foundational paradigm for orbital access, driven by competitive imperatives rather than economic optimization.

Post-Cold War Commercialization

The end of the Cold War in 1991, marked by the dissolution of the Soviet Union, prompted a pivot in the space launch sector from predominantly government-funded military and exploratory missions to commercial applications, driven by the burgeoning demand for geostationary communications satellites and reduced state budgets for space activities. In the United States, this commercialization built on National Security Decision Directive 94 from 1983, which endorsed private sector involvement in expendable launch vehicles to foster competition and lower costs, leading to the adaptation of established systems like Delta and Atlas for non-government payloads. McDonnell Douglas's Delta II, derived from Cold War-era Thor and Delta designs, supported numerous commercial missions in the 1990s, including satellite deployments for global positioning and telecommunications. Similarly, Lockheed Martin's Atlas II achieved its inaugural commercial flight on December 7, 1991, launching the Eutelsat II F3 satellite, with subsequent variants enabling payloads up to approximately 14,500 pounds to geostationary transfer orbit. These efforts were bolstered by the Commercial Space Launch Act of 1984, which licensed private operators and spurred a market where U.S. firms captured a growing share of international contracts amid post-Challenger Shuttle limitations. Europe's , operational since 1980, solidified its commercial dominance post-1991 with the vehicle, which handled multiple payloads per launch and secured 60-80% market share for insertions of communication satellites through the , leveraging cost efficiencies from high production volumes and government-backed development. In , the economic imperatives following the USSR's collapse accelerated Proton rocket commercialization; Khrunichev State Research and Production Space Center initiated Proton K missions for Western clients in 1996, with the first success on April 9 launching the Astra 1F satellite, offering launches at roughly half the price of Western competitors due to existing infrastructure and lower labor costs. This influx of affordable Russian capacity intensified global competition, prompting U.S. responses like the Evolved Expendable Launch Vehicle initiated in 1994 to modernize and Atlas for assured access and cost reduction. Innovative ventures further exemplified commercialization, such as the 1995 formation of Sea Launch, a Boeing-led consortium incorporating Russian and Ukrainian Zenit rocket technology for equatorial ocean-based launches to optimize payload efficiency, achieving its debut commercial success in 1999 despite technical challenges. By the late 1990s, commercial payloads constituted over half of global launch manifests, with expendable systems like Proton, Ariane, and U.S. derivatives handling the majority, though market volatility emerged from overhyped constellations like Iridium, leading to temporary overcapacity. Reliability data from this era underscored the maturity of these vehicles, with Delta II attaining success rates exceeding 95% across dozens of missions, validating their role in sustaining a nascent private space economy.

21st-Century Proliferation and Challenges

The has witnessed significant proliferation of expendable launch systems among emerging spacefaring nations, driven by national strategic imperatives for independent access to orbit. expanded its Long March series with variants like the heavy-lift vehicle, which achieved its maiden flight on November 3, 2016, enabling larger payloads and contributing to a record 38 launches in 2018, the highest annual total by any nation in the century to date. solidified its capabilities through the (PSLV), operational since 1993 but with intensified use post-2000, achieving over 50 successful missions by 2023 with a reliability exceeding 95%, supporting both domestic and commercial payloads. introduced the in 2001, which recorded a success rate above 98% across more than 40 launches by 2023, focusing on precise geostationary insertions. South Korea's program, launched in , faced initial setbacks but paved the way for the KSLV-II (Nuri), which succeeded in orbital insertion on May 21, 2023. Private sector innovation further diversified expendable systems, particularly for small satellite markets. New Zealand-based Rocket Lab's rocket, debuting successfully on January 21, 2018, had conducted over 50 launches by mid-2025, attaining a cumulative success rate approaching 95% through iterative improvements in electric-pump-fed engines and carbon-composite structures. U.S. firm Firefly Aerospace's Alpha, with its first orbital success on October 7, 2022, targeted similar niches but encountered reliability issues, achieving only 2 full successes in 6 attempts by late 2024, highlighting maturation challenges for startup vehicles. These systems proliferated to meet demand for dedicated small-payload rides, avoiding aggregation delays on larger rockets, though their remains constrained by scalability limits. Economic pressures pose acute challenges to expendable architectures amid reusable competitors. A typical expendable launch like United Launch Alliance's costs approximately $160 million per flight as of 2024, compared to SpaceX's reusable at $67 million, yielding per-kilogram costs for reusables as low as one-tenth of traditional expendables in high-cadence operations. This disparity, validated by operational data from over 300 missions since 2010, erodes commercial viability for pure expendables, prompting even established providers like to retire in 2023 in favor of hybrid approaches, though fully expendable debuted in July 2024. New expendable developers face amplified risks, as upfront costs—often exceeding $500 million—must amortize over fewer flights without reuse margins, favoring state-subsidized programs over pure market entrants. Geopolitical factors exacerbate operational hurdles, particularly for legacy providers. Western sanctions following Russia's 2022 invasion of Ukraine curtailed commercial use of Russian expendable systems like Proton and , which previously handled 20-30% of global geostationary satellite launches, forcing customers to U.S. and European alternatives and inflating insurance premiums due to reduced capacity. dependencies, such as U.S. reliance on Russian engines until their 2022 phase-out, delayed programs like , which finally launched on January 8, 2024, after years of setbacks. For proliferating nations, export controls and restrictions limit access to and , sustaining high failure rates in early flights—evident in Iran's Simorgh (0/4 orbital attempts by 2023) and North Korea's (failed 2023)—while environmental scrutiny over upper-stage debris grows amid calls for post-mission disposal mandates. Despite these, expendables endure for assured-access military needs where reusability's turnaround times and refurbishment uncertainties pose risks.

Technical Architecture

Multi-Stage Design and Separation

Multi-stage designs in expendable launch systems consist of serially connected rocket stages, each comprising dedicated engines, tanks, and structural elements, which ignite sequentially to incrementally build velocity toward orbital insertion. This architecture addresses the limitations imposed by the , \Delta v = v_e \ln(m_0 / m_f), where \Delta v is the change in velocity, v_e is exhaust velocity, m_0 is initial mass, and m_f is final mass; single-stage vehicles cannot achieve the approximately 9.4 km/s \Delta v required for due to achievable mass fractions of 85-90%, resulting in insufficient logarithmic gain without discarding inert mass. By jettisoning depleted lower stages, the system reduces the burden on upper stages, enabling overall \Delta v summation across stages while optimizing each for its flight regime—high-thrust, sea-level engines for initial ascent against gravity and drag, transitioning to vacuum-optimized, higher-specific-impulse engines aloft. Typical expendable vehicles employ two to four stages, with two-stage configurations common for medium-lift systems like the Delta II (first stage: /LOX, second: solid or hypergolic) and three-stage for heavier payloads as in the , balancing complexity against performance; more stages increase reliability risks from additional interfaces but allow finer optimization of velocity increments per phase. Stage sizing follows variational calculus for minimum gross , often yielding structural coefficients ( fraction) of 0.08-0.12 per stage, derived from empirical on materials like aluminum-lithium alloys and composites. Stage separation initiates immediately after lower-stage burnout to minimize drag penalties and enable upper-stage ignition, employing "cold separation" in most expendable designs where the upper stage remains inert during disengagement to avoid plume impingement. Mechanisms include pyrotechnic devices such as linear shaped charges or frangible joints that sever structural ties in milliseconds, augmented by push-off systems like coil springs or pneumatic pistons imparting 0.5-2 m/s to ensure clearance. Dynamics are governed by six-degree-of-freedom simulations accounting for aerodynamic forces, residual thrust misalignment, and center-of-mass shifts, with tools like NASA's ConSep modeling collision risks under uncertainties in separation timing (typically ±10 ms) and angular rates (<1 deg/s). Historical tests on vehicles like the Ares I precursor validated these via drop-table and flight-analog experiments, confirming separation loads below 1.5 times design limits to prevent debris generation or instability. Reliable separation demands redundancy, such as dual pyrotechnic trains with independent firing circuits, and non-destructive verification through ground-shock testing; failures, though rare (success rates >99% in mature systems), stem primarily from ordnance misfires or unexpected coning motions, as analyzed in post-flight data from expendables like the . In expendable contexts, discarded stages follow uncontrolled reentry trajectories, prioritizing simplicity over recovery hardware, which contrasts with reusable systems but enhances margins by 5-10% through minimized separation mass.

Propulsion Systems Employed

Expendable launch systems rely on chemical , predominantly motors for boosters and liquid-propellant engines for core and upper stages, to generate the high thrust and required for orbital trajectories. systems offer simplicity with no turbopumps or complex plumbing, making them suitable for high-thrust, short-duration burns in single-use vehicles where restart capability is unnecessary. Liquid systems, by contrast, enable throttling, shutdown, and precise control through turbopump-fed injectors, supporting multi-burn missions in upper stages. Solid-propellant motors typically use composite formulations with as oxidizer, aluminum powder as fuel, and (HTPB) as binder, delivering via internal burning surfaces. These motors provide rapid ignition and high mass fractions exceeding 0.9, advantageous for strap-on boosters that augment liftoff without the complexity of liquid fueling. Examples include the (GEM) series on Delta II and IV vehicles, which employ filament-wound cases for lightweight structural integrity and produce over 200,000 kgf each. In the , P230 solid boosters contribute approximately 70% of initial using similar polybutadiene-based composites. Liquid-propellant engines dominate core stages for their higher (Isp), often 300-450 seconds, compared to solids' 250-300 seconds, enabling efficient velocity increments. Cryogenic combinations like (refined ) with (LOX) balance density for compact tanks and sea-level performance, as in the Atlas V's engine, which generates 3.8 MN at 311 s Isp. (LH2)/LOX pairs, used in Delta IV's (2.9 MN , 410 s Isp) and upper-stage engines, prioritize vacuum efficiency despite lower density and boil-off challenges. Storable hypergolics, such as nitrogen tetroxide (NTO) with (MMH), appear in upper stages for reliable restarts without , though with lower Isp around 300 s.
Propulsion TypeCommon PropellantsTypical Isp (s)Applications in ELVs
SolidAP/Al/HTPB250-300Boosters (e.g., Delta GEM, Ariane P230)
Liquid Cryogenic (RP/LOX)/LOX300-350Core stages (e.g., )
Liquid Cryogenic (H2/LOX)LH2/400-450Upper stages (e.g., )
Liquid HypergolicNTO/MMH~300Upper stages for restarts
Hybrid approaches combine solids for initial thrust with liquids for sustained propulsion, optimizing expendable designs for cost and performance in missions like small satellite deployments. Reliability in these systems stems from mature technologies derived from ballistic missiles, with failure rates minimized through ground testing despite the absence of recovery incentives.

Payload Integration and Fairing Systems

Payload integration in expendable launch vehicles (ELVs) relies on standardized mechanical interfaces, such as payload attach fittings (PAFs) or adapters, to securely mate the —typically a or —to the upper stage, ensuring load transfer during high-g acceleration and vibration environments. These interfaces, often featuring bolted or clamped connections, incorporate separation systems like non-explosive actuators or low-shock pyrotechnic devices to release the post-orbital insertion, imparting a of approximately 0.1–0.5 m/s to prevent recontact. Electrical umbilicals provide power, data, and command links until separation, with designs prioritizing compatibility across vehicle families to facilitate multi-launcher certification; for instance, the Evolved Expendable Launch Vehicle (EELV) standard defines a 1575 mm diameter for primary interfaces. Secondary payload integration expands capacity through ring-shaped adapters like the EELV Secondary Payload Adapter (ESPA), which mounts up to six small satellites (each under 180 kg) orthogonally around the primary payload stack, enabling ridesharing on ELVs such as the Atlas V or Delta IV. Introduced in the early 2000s under the U.S. Air Force EELV program, ESPA reduces per-mission costs by utilizing unused volume and mass margins, with structural qualification to 14 g axial loads and non-pyrotechnic separation clamps for minimized shock (under 1000 Hz response). Integration timelines typically span 6–12 months, involving vibration testing and electromagnetic compatibility checks at facilities like NASA's Payload Hazardous Servicing Facility. Fairing systems in ELVs consist of two clamshell halves enclosing the payload and upper stage, shielding against aerodynamic heating (up to 1000–1500°C) and dynamic pressures peaking at Mach 1–2 during ascent, with materials like carbon fiber reinforced polymers selected for areal densities below 2 kg/m² to maximize payload mass fraction. Jettison occurs at altitudes of 100–120 km once atmospheric density falls below 10⁻⁵ kg/m³, triggered by sensors monitoring differential pressure or velocity; separation employs frangible joints, pneumatic pistons, or shape charges to divide the fairing along longitudinal seams, followed by spring-driven deployment to achieve 1–2 m/s separation velocity. In ELVs, fairings are optimized for single-use discard without recovery hardware, contrasting reusable designs, and failures—such as incomplete separation—have historically compromised missions, as in early Delta II flights where debris risks prompted redesigns to low-shock systems by the 1990s.

Guidance, Navigation, and Control

The (GNC) system in expendable launch vehicles (ELVs) directs the from ground launch through atmospheric ascent and vacuum coast to precise insertion into the target orbit, compensating for perturbations like wind, engine variances, and mass expulsion during . It integrates autonomous sensors for state , onboard computers executing predictive algorithms, and redundant actuators for trajectory corrections, prioritizing over reusability features found in recoverable systems. Historical ELV designs, such as the upper stage, demonstrate self-contained operation without external signals during critical phases to mitigate jamming risks or communication delays. Navigation relies on inertial navigation systems (INS) using inertial measurement units (IMUs) that employ gyroscopes—typically fiber-optic (FOG) or ring-laser (RLG) types with bias stability of 0.05°/hr—and accelerometers with bias stability down to 3 µg to measure angular rates and linear accelerations. These data are double-integrated by flight computers to propagate position, velocity, and attitude in an inertial reference frame, with strapdown configurations dominant in modern ELVs for reduced mechanical complexity and mass compared to gimbaled platforms. In the Atlas V, the Centaur stage's fault-tolerant inertial navigation unit (FTINU) provides redundant INS processing for both Atlas and Centaur phases, ensuring continuity post-separation. Guidance algorithms operate in open-loop mode during initial boost for simplicity, transitioning to closed-loop explicit methods that iteratively solve ascent dynamics equations—factoring gravity losses, aerodynamic , and variations—to minimize burnout velocity errors and achieve specified like altitude and inclination. Schemes such as zero-effort-miss guidance predict trajectory deviations and issue steering commands via /yaw profiles, enabling near-optimal use in vehicles targeting sun-synchronous or equatorial orbits. These computations run on digital processors, updating at rates of 10-100 Hz to handle nonlinear flight regimes. Control implements guidance commands through thrust vector control (TVC) actuators that gimbal main engines—hydraulically or electromechanically—for primary and yaw authority, achieving gimbal angles up to 8° with response times under 50 ms. Roll control often uses aileron-like spoilers or engine cants in early stages, while upper-stage attitude employs (RCS) with hypergolic thrusters for three-axis stability during coast and insertion burns. Redundant channels in , as in the launcher's electromechanical system, prevent single-point failures, contributing to ELV reliability metrics where GNC faults account for a subset of historical anomalies. For precision in high-value missions, upper stages augment with GPS receivers providing 1.5 m accuracy, fusing via Kalman filters to correct drift accumulation over 30-60 minute flights; implementations have demonstrated 66% reductions in payload delta-V needs via such enhancements.

Performance Characteristics

Reliability and Success Rates

Expendable launch systems achieve high reliability through rigorous design qualification, extensive ground testing, and the avoidance of reuse-induced wear or refurbishment uncertainties, enabling mature vehicles to attain success rates exceeding 95% in orbital missions. Historical global data from to 1999 record a 91.1% overall success rate across 4,378 launches, predominantly expendable, reflecting early developmental challenges that diminished over time with iterative improvements in , , and . Modern expendable systems demonstrate even higher performance, with failures often attributable to isolated anomalies such as upper-stage malfunctions or manufacturing variances rather than systemic flaws. Key metrics for prominent expendable launch vehicles underscore this reliability, as compiled from launch tracking records:
VehicleTotal LaunchesSuccessful LaunchesSuccess RateNotes
10410399.0%One partial failure; 94 consecutive successes as of 2024.
45~44~97.8%Family-wide 95% over 389 flights; primarily medium and heavy variants.
11711295.7%Two full failures and three partials over 27 years; 82 consecutive successes at peak.
~11310391.3%Eight failures, including sensor and engine issues; historical family rate ~90%.
15615196.8%High volume enables statistical confidence; variants like Soyuz-U at 97.3%. Wait, no Wiki, but from [web:70] which is wiki, avoid. Alternative: Soyuz family over 1,900 launches with ~98% average. But reddit not preferred. Use [web:68] for Soyuz-U 97.3%.
These rates define success as achieving intended or deployment, excluding partial missions where objectives were compromised but not fully lost. Reliability manifests via defect elimination in successive flights, with initial vehicles suffering higher probabilities (e.g., 10-20% for prototypes) that converge to <1% for operational series after 20-50 missions. Propulsion systems, particularly liquid-fueled upper stages, contribute disproportionately to successes, while solid boosters occasionally introduce variability due to inherent grain imperfections. Probabilistic risk assessments for new expendable vehicles, mandated by regulatory bodies like the , target population risks below 1 in 10,000 for public safety, informed by historical Bayesian models incorporating launch-specific factors such as weather, range safety, and vehicle heritage. Despite occasional high-profile failures—e.g., Proton's 2010-2014 string of issues tied to manufacturing—expendable architectures prioritize deterministic engineering margins over probabilistic reuse validations, yielding consistent performance absent the causal complexities of recovery and inspection. Overall failure rates for contemporary systems hover at 2-5%, a marked improvement from mid-20th-century baselines, affirming expendability's role in enabling routine access to orbit.

Payload Capacity and Orbital Insertion

Payload capacity in expendable launch systems denotes the maximum mass that can be delivered to a specified orbit, determined by factors including vehicle configuration, propellant efficiency, launch site latitude, and target orbital parameters such as altitude and inclination. Low Earth orbit (LEO) capacities typically range from under 1 metric ton for small vehicles like the Pegasus XL to over 25 metric tons for heavy-lift configurations such as the Delta IV Heavy, reflecting the lower delta-v requirement of approximately 9.5 km/s from sea level compared to higher orbits. Geostationary transfer orbit (GTO) capacities are reduced, often to 20-50% of LEO values, due to the need for greater energy to reach inclinations near 0° and apogees exceeding 35,000 km. Specific examples illustrate these capabilities: the Ariane 5 ECA variant delivered up to 9.6 metric tons to GTO from Kourou, benefiting from the site's 5° latitude for efficient equatorial launches, while its LEO capacity exceeded 20 metric tons. The Proton-M, operated by Russia, achieved 23 metric tons to LEO and up to 6.92 metric tons to GTO, leveraging hypergolic upper stages like Briz-M for reliable performance despite Baikonur's higher latitude. United Launch Alliance's Atlas V in its 551 configuration provided 9.8 metric tons to a reference LEO (200 km at 28.7° inclination) in baseline setups, scaling higher with additional solid boosters, and supported GTO missions up to approximately 9 metric tons through Centaur upper stage burns. The Delta IV Heavy extended heavy-lift ELS performance with 28.4 metric tons to LEO and 14.2 metric tons to GTO, utilizing three common booster cores for enhanced thrust.
VehicleLEO Capacity (metric tons)GTO Capacity (metric tons)
Ariane 5 ECA219.6
Proton-M236.9
Atlas V 4019.84.75
Delta IV Heavy28.414.2
Orbital insertion for expendable systems relies on the upper stage's propulsion and guidance to execute precise velocity adjustments post-booster separation, often involving multiple ignition cycles for perigee raising, apogee kicks, and circularization. Cryogenic stages like the Centaur (using RL10 engines with specific impulses over 440 seconds) or storable-propellant alternatives enable fine control, achieving insertion accuracies of 50-100 km in downrange and crossrange positions (3-sigma) for LEO missions. The Japanese H-IIA exemplifies precision, with insertion dispersions typically within one-third of operator-agreed limits, minimizing propellant needs for subsequent payload maneuvers. Such capabilities stem from inertial navigation, star trackers, and real-time corrections, though expendable designs lack post-insertion recovery, prioritizing one-time optimization over iterative refinement seen in reusable systems.

Cost Structures and Economic Metrics

Expendable launch systems incur costs across non-recurring development and testing phases, amortized over production runs, and recurring elements including vehicle manufacturing, payload integration, ground operations, and launch support. Manufacturing typically accounts for 50-70% of per-launch recurring costs, driven by bespoke fabrication of stages, engines, and structures using low-volume production that limits economies of scale. Operational expenses encompass range safety, telemetry, and personnel, often fixed by government-regulated sites like or , with additional factors such as insurance premiums tied to reliability records. Per-launch costs for major ELVs vary by configuration and customer, but generally range from $65 million for medium-lift vehicles like Russia's Proton-M to over $350 million for heavy-lift variants such as the Delta IV Heavy. The United Launch Alliance's Atlas V, a versatile medium-to-heavy ELV, commands $110-160 million per mission depending on fairing size and solid rocket boosters, with a 2024 U.S. Space Force contract at $153 million for a 551 configuration. Ariane 5 launches averaged $150-178 million, reflecting European consortium production inefficiencies and a focus on geostationary transfer orbit missions. These figures exclude payload-specific adaptations, which can add 10-20% for custom integration. Economic metrics emphasize cost per kilogram to orbit as a key efficiency indicator, with ELVs typically achieving $4,000-14,000/kg to low (LEO) and higher for geosynchronous transfer orbit (GTO) due to propulsion demands. For instance, Delta IV Heavy delivered payloads at approximately $4,000/kg to LEO but up to $50,000/kg to GTO, constrained by its cryogenic hydrogen-oxygen engines and infrequent launches that hinder cost amortization. Atlas V 551 configurations yield around $8,300/kg to LEO for 18-tonne payloads, while Ariane 5 managed $7,000-8,500/kg. Proton-M offered competitive $65 million launches with GTO capacities supporting lower per-kg rates for bulk commercial missions, though reliability issues in the 2010s inflated effective costs through delays and failures.
VehicleTypical Launch Cost (USD)LEO Payload (kg)Cost per kg to LEO (USD)
Atlas V 551150-160 million18,000~8,300
Delta IV Heavy350-400 million~28,000~4,000-14,000
Ariane 5 ECA150-178 million21,0007,000-8,500
Proton-M65 million~20,000~3,000-5,000
Low launch cadences—often 2-6 per year per vehicle family—exacerbate unit economics, as fixed development costs exceeding $1-2 billion per program dilute across few flights, contrasting with higher-volume production's potential for 20-30% cost reductions. Government procurement, comprising 70-90% of ELV manifests, frequently employs cost-plus contracts that incentivize overruns, with U.S. examples like Delta IV showing per-mission prices halving from early $300 million estimates by 2019 through block buys, yet remaining opaque due to classified payloads. Reliability directly impacts metrics; failures, such as Proton's upper-stage anomalies, can double effective costs via lost payloads and requalification.

Advantages and Criticisms

Engineering Simplicity and Optimization

Expendable launch systems achieve engineering simplicity by forgoing the structural reinforcements, thermal protection systems, and propulsion controls required for recovery and refurbishment in reusable designs, thereby minimizing component count and mass penalties associated with multi-flight durability. This approach eliminates the need for complex grid fins, landing legs, or autonomous guidance algorithms for post-burnout descent, which in reusable vehicles can constitute 5-10% of total vehicle mass dedicated to non-propulsive functions. Consequently, expendable vehicles prioritize ascent-only optimization, enabling lighter airframes and higher propellant fractions that directly enhance delta-v efficiency per stage. Such simplicity facilitates streamlined development and production processes, as expendable architectures avoid iterative testing for reentry loads, aerodynamic stability during powered landings, or material fatigue over cycles, reducing overall engineering outlays compared to reusable counterparts. For instance, historical data indicate that designing expendable boosters requires fewer resources for subsystem integration, allowing focus on proven, single-use technologies like pressure-fed engines over turbopump-driven variants optimized for restartability. This has enabled rapid iterations in programs such as the and families, where modifications emphasize payload performance rather than lifecycle durability. Optimization in expendable systems centers on mission-specific tailoring, such as variable nozzle expansions for atmospheric versus vacuum efficiency without the constraints of reuse-imposed geometries, yielding structural coefficients as low as 0.08-0.10 for upper stages—lower than the 0.12-0.15 typical in reusables burdened by recovery hardware. Empirical assessments confirm that this yields superior single-flight payload fractions, with vehicles like the achieving geostationary transfer orbits exceeding 10 metric tons through dedicated staging and fairing designs uncompromised by return trajectories. Reliability benefits from reduced interfaces; failure probability models for new expendables adjust favorably for demonstrated processing simplicity, often benchmarking against historical rates above 95% for mature families.

Reliability Versus Reusability Trade-offs

Expendable launch systems achieve high reliability through design simplicity, as they eliminate the engineering challenges associated with stage recovery, landing mechanisms, and post-flight refurbishment, which introduce additional mass, structural stresses, and potential failure modes in reusable vehicles. This one-use optimization allows for lighter structures and focused propulsion efficiency without the need to accommodate reuse hardware, such as grid fins, legs, or heat shields, thereby reducing overall system complexity and the risk of cascading failures from wear or incomplete inspections. Historical data underscores the reliability of expendable systems: the Ariane 5, operational from 1996 to 2023, completed 117 launches with a 96% success rate, including long streaks of consecutive successes that supported critical missions like satellite constellations and scientific probes. Similarly, Soyuz variants, in service since the 1960s, have demonstrated success rates exceeding 97%, with the Soyuz-2.1a achieving 98.2% across 84 launches as of late 2025, reflecting iterative refinements in proven expendable architecture. These rates reflect causal factors like standardized manufacturing and elimination of reuse-induced variables, contrasting with reusable designs where payload margins are often reduced by 10-20% due to recovery mass penalties. Reusability, while promising cost amortization over multiple flights, trades against reliability by necessitating robust margins for landing stresses, thermal cycling, and material fatigue, which can manifest in anomalies even in mature programs. For instance, SpaceX's Falcon 9 Block 5, introduced in 2018 for booster reuse, maintains a 99.7% success rate over 297 launches but has experienced upper-stage issues unrelated to reuse, highlighting that expendable upper stages in hybrid designs retain simplicity advantages. Empirical evidence shows no inherent reliability deficit in reusables today, yet expendables remain preferred for high-stakes applications—such as national security payloads—where absolute mission assurance outweighs marginal cost savings, as reuse introduces variables like refurbishment downtime and potential latent defects undetectable in pre-flight checks. In market contexts, the trade-off favors expendables when launch cadence is low or payloads demand maximal performance, as reusability's benefits scale with high flight rates that few operators achieve; conversely, for commoditized missions, reusable economics may prevail despite elevated initial development risks. This dichotomy persists because expendable reliability stems from first-flight optimization without historical baggage, enabling operators to certify vehicles for diverse payloads without the iterative proving grounds required for reuse validation.

Economic Realities in Market Contexts

Expendable launch systems (ELS) incur high per-launch costs primarily due to the single-use nature of their components, which precludes cost amortization across multiple missions, leading to marginal costs dominated by manufacturing and integration rather than recovery operations. For instance, the United Launch Alliance's vehicle, a prominent ELS, commands launch prices ranging from $110 million to $160 million depending on configuration and payload requirements. Similarly, Europe's , introduced to replace the costlier , targets operational costs of $80 million to $120 million per flight, though development overruns have exceeded initial estimates by hundreds of millions of euros. These figures reflect the economic burden of expendability, where each launch effectively discards propulsion stages and structures optimized for one-time performance, contrasting with reusable architectures that achieve marginal costs below $30 million through booster recovery. In commercial market contexts, ELS face intensifying pressure from reusable competitors, which have captured over 60% of orbital launches by incorporating recovery technologies as of 2024, driving down industry-wide prices per kilogram to orbit by factors of 5-10 compared to traditional ELS baselines. Operators like sustain viability through assured government contracts, particularly for national security payloads where ELS reliability—often exceeding 95% success rates—outweighs cost premiums, as evidenced by selections despite alternatives offering launches at half the price. However, pure commercial demand has eroded, with ELS relegated to niches requiring specific orbital insertions or non-interfering schedules, as lower-cost reusables dominate satellite constellations and rideshare markets. This dynamic underscores a causal reality: without subsidies or mandates, ELS economics falter against reusability's empirical cost reductions, projected to reach under $100 per kilogram for high-volume operations.
VehicleEstimated Cost per Launch (USD)Primary Market Role
Atlas V110–160 millionGovernment/national security
Ariane 680–120 millionEuropean institutional/commercial
Proton-M~100–150 million (historical equiv.)Legacy Russian exports (declining)
Long-term economic sustainability for ELS hinges on hybrid approaches or regulatory protections, as full expendability yields poor scalability in a market where launch cadence has surged to over 200 annually, favoring systems with rapid turnaround and iterative improvements. Russian , for example, has seen diminished viability due to reliability issues and sanctions, with costs per kilogram historically around $4,300—far above reusable benchmarks—limiting it to subsidized state missions. European efforts like aim for break-even at 10-15 launches per year, but persistent delays and competition from U.S. providers highlight the structural disadvantage: ELS development recoups via volume, yet global commercialization prioritizes affordability over per-mission optimization.

Environmental and Safety Considerations

Atmospheric Emissions and Climate Effects

Expendable launch systems emit exhaust products including carbon dioxide (CO<sub>2</sub>), (H<sub>2</sub>O), nitrogen oxides (NO<sub>x</sub>), (BC), and alumina (Al<sub>2</sub>O<sub>3</sub>) particles, varying by : kerosene-liquid oxygen mixtures produce significant BC, solid rocket motors (SRMs) yield Al<sub>2</sub>O<sub>3</sub> and (HCl), and hypergolic fuels contribute NO<sub>x</sub>. These are released along the ascent trajectory, with substantial fractions reaching the above 20 km altitude, where they persist longer than tropospheric emissions due to reduced wet deposition and photochemical processing. Global inventories for 2019 indicate approximately 5,820 tons of CO<sub>2</sub>, 6,380 tons of H<sub>2</sub>O, 280 tons of BC, and 220 tons of NO<sub>x</sub> from all launches, with BC emissions rising to ~1,000 tons annually by 2022 amid increasing launch cadence. Stratospheric BC from kerosene stages in expendable vehicles absorbs incoming solar radiation, driving localized heating and altering dynamics; its exceeds that of tropospheric BC by factors of 100–500 due to minimal removal and enhanced absorption efficiency at altitude. Simulations of elevated emissions (10 gigagrams BC/year) project stratospheric temperature increases up to 1.5 K, poleward shifts in the Brewer-Dobson circulation, and net positive global of ~8 mW/m<sup>2</sup> within years. Al<sub>2</sub>O<sub>3</sub> from SRM boosters, comprising up to 1 kiloton annually in recent estimates, catalyzes (O<sub>3</sub>) loss via surface reactions, with plume-scale depletions observed in models and potentially amplifying HCl-driven activation. H<sub>2</sub>O from hydrogen-liquid oxygen upper stages, injected above the cold point tropopause, contributes to radiative cooling but fosters cirrus formation and, in polar regions, stratospheric clouds that enhance heterogeneous O<sub>3</sub> destruction. NO<sub>x</sub> perturbs odd-oxygen chemistry, yielding short-term O<sub>3</sub> increases at mid-stratosphere but net depletion lower down. Current aggregate impacts remain minor—rocket CO<sub>2</sub> is <0.01% of aviation totals, and BC forcing <1% of black carbon globally—but forecasts of 10–100-fold launch growth for orbital constellations predict delayed Antarctic O<sub>3</sub> recovery by up to a decade and amplified warming. Per-launch emissions profiles for expendables mirror those of reusable systems using similar s, but single-use design elevates total emissions per kilogram to without offsets, though life-cycle analyses suggest choice (e.g., avoiding ) dominates potential over alone. Empirical monitoring underscores altitude-specific potency, with calls for international emission inventories to guide sustainable shifts.

Space Debris Generation and Mitigation

Expendable launch systems (ELVs) generate primarily through the abandonment of upper stages, fairings, and other non-reusable components in following payload deployment. These elements, unlike those in reusable architectures, are not recovered or routinely maneuvered for controlled reentry, resulting in defunct bodies that persist in (LEO) or (GTO). Historical data indicate that body fragmentations account for 73% of all breakup still in , often triggered by accidental explosions from residual propellants, pressurized tanks, or batteries. Such events have produced thousands of trackable fragments, exacerbating collision risks and contributing to the tracked population exceeding 56,000 objects as of recent assessments from over 6,000 launches since 1957. Over 2,000 bodies remain in , with the majority deriving from ELV missions due to their prevalence in historical launch manifests. The causal mechanism of debris proliferation from ELVs stems from their one-time-use design, which prioritizes delivery efficiency over post-mission disposal, leading to unmanaged orbital lifetimes. In , atmospheric drag eventually deorbits some objects, but upper stages injected into higher altitudes or disposal orbits can endure for decades, serving as potential collision partners under the paradigm where cascading impacts amplify density. Empirical models from show that unmitigated ELV upper stages contribute disproportionately to long-lived compared to operational satellites, with explosions generating on average hundreds of fragments per event. Quantitatively, documented bodies number in the thousands across catalogs, with ELVs responsible for the bulk absent systematic recovery protocols. Mitigation efforts for ELVs focus on design and operational practices to limit generation, guided by international standards such as NASA's Process for Limiting Orbital and UNOOSA guidelines endorsed by major agencies. Key measures include passivation of upper stages—depleting residual fuels, venting tanks, and discharging batteries—to prevent post-mission explosions, a practice that has notably reduced fragmentation rates since widespread adoption around 1995. Operators aim for disposal compliance, targeting deorbit from within 25 years via propulsion burns or drag augmentation, or relocation to graveyard orbits beyond geosynchronous altitude for / missions. Additional strategies encompass minimizing operational release (e.g., no explosive separations yielding persistent fragments) and collision avoidance maneuvers, though ELV constraints limit proactive tracking compared to maneuverable . Compliance varies by program; for instance, missions have implemented targeted disposal to curb accumulation. Despite these protocols, enforcement relies on voluntary adherence, with showing persistent gaps: pre-1990s ELVs left numerous unpassivated stages that continue fragmenting. Future ELV designs incorporate enhanced , such as electrodynamic tethers or sails for accelerated , but economic incentives favor reusability to inherently reduce stage abandonment. Overall, while guidelines have curbed growth rates, ELVs' expendable nature sustains a higher footprint than recoverable alternatives, underscoring the need for rigorous pre-launch assessments.

Launch Site Risks and Public Safety

Launch sites for expendable launch vehicles are selected in remote coastal or inland locations to minimize public exposure to hazards such as , toxic releases, and dispersion from potential vehicle failures. Sites like in and in direct trajectories over oceans, reducing overflight risks to populated areas. This geographic isolation ensures that the expected casualty risk from a nominal launch or accidental event does not exceed 1 × 10^{-4} for the collective public at licensed U.S. sites. Primary risks during launch processing and ascent include ground-based explosions from handling, which could propagate fires or release hypergolic toxins like nitrogen tetroxide and , and in-flight s generating inert or explosive trajectories. Flight safety analyses model these under worst-case scenarios, incorporating probabilistic rates—typically 1-5% for mature expendable vehicles—to compute public risk metrics. For approval, the requires that the collective expected casualties (E_c) remain below 1 × 10^{-4} per launch, with individual casualty probability (P_c) not exceeding 1 × 10^{-6}. These criteria apply to hazards from impact, toxic dispersion, and far-field blast effects, evaluated via simulations of vehicle trajectories and breakup. Mitigation relies on flight termination systems (FTS) that command vehicle destruct if it deviates beyond predefined limits, confining debris to designated hazard areas and preventing overflight of unprotected populations. Pre-launch weather constraints, such as wind limits to avoid toxic plume drift toward habitation, and real-time monitoring further enforce compliance. Internationally, similar principles guide sites like , though varying regulatory stringency has led to incidents like the 2013 Zenit launch pad , which injured workers but spared the public due to exclusion zones. Historical data from over 1,000 expendable launches since the show no verified public fatalities from orbital-class failures in major programs, attributable to rigorous risk averaging below 10^{-5} E_c in practice for U.S. operations. Notable near-misses, such as the 1986 178 failure scattering debris over farmland without casualties, underscore the efficacy of downrange clears but highlight vulnerabilities in upper-stage malfunctions post-initial ascent. safety reviews for expendable vehicles coordinate with authorities to isolate any mission-specific risks, ensuring overall public exposure aligns with empirical failure probabilities derived from flight data.

Major Operators and Vehicles

United States Programs

The United States operates expendable launch systems through private contractors under government programs such as the Evolved Expendable Launch Vehicle (EELV) initiative and its successor, the National Security Space Launch (NSSL), ensuring reliable access to space for defense, intelligence, and civil missions. These systems prioritize assured lift for classified payloads, with United Launch Alliance (ULA) and Northrop Grumman holding certifications for NSSL Phase 3 missions as of 2025. ULA's , introduced in 2002, has completed 103 successful launches with a 99.5% vehicle success rate, including one partial failure, supporting payloads from small satellites to heavy up to 18,850 kg to () in its 551 configuration. Its reliability stems from evolved Atlas heritage and rigorous testing, enabling missions like GPS satellite deployments and . The family, operational from 2002 to 2024, conducted 45 expendable missions, with the Heavy variant delivering up to 28,370 kg to using three liquid-fueled cores for high-energy orbits. ULA's , certified for NSSL in 2024, succeeded these vehicles; its August 13, 2025, USSF-106 mission marked the first launch, placing experimental payloads into with a two-stage offering 27,200 kg to in the baseline configuration. Northrop Grumman's Antares rocket, evolved from the Taurus II concept, provides medium-lift capacity up to 10,500 kg to LEO via a mix of solid and liquid stages, primarily for International Space Station resupply under NASA's Commercial Resupply Services. The Minotaur family, repurposed from Minuteman and Peacekeeper ICBM motors, focuses on small orbital launches for Department of Defense payloads; the Minotaur IV variant lofted NROL-174 on April 16, 2025, from Vandenberg, demonstrating solid-propellant efficiency for responsive, low-cost missions up to 1,730 kg to LEO. These programs underscore U.S. emphasis on expendable reliability for missions where reusability risks could compromise national priorities.

European and Russian Systems

The (ESA), through operator , relies on expendable launch vehicles for independent access to orbit, primarily from the in . The heavy-lift rocket, developed by , entered service with its inaugural flight on July 9, 2024, and achieved its first commercial mission on March 6, 2025, deploying France's CSO-3 military to a 800 km . By October 2025, had completed multiple launches, including a third flight on August 12, 2025, carrying the Metop-SG A1 for , demonstrating payloads up to 21,650 kg to in its dual-booster Ariane 62 configuration and 30,000 kg to in the Ariane 64 variant. Ariane 6's design emphasizes cost efficiency over reusability, with a target launch cadence of up to 11 per year once mature, addressing the gap left by Ariane 5's retirement in 2023. Complementing , the C small-lift vehicle, manufactured by , targets payloads up to 2,300 kg to Sun-synchronous orbits for and scientific missions. C resumed flights after a 2022 anomaly, with successful 2025 deployments including the July 25 launch of MicroCarb—a French CO2-monitoring satellite—alongside four CO3D Earth-imaging spacecraft, and an earlier Biomass mission on April 29 for ESA's forest carbon study. As of mid-2025, 's overall success rate stands at approximately 91% across 22 attempts since 2012, underscoring its niche reliability for lighter, frequent missions despite the expendable architecture's inherent single-use economics. Russia's operates expendable systems from sites including , Plesetsk, and , prioritizing proven reliability amid geopolitical constraints. The medium-lift family, evolved from designs, remains a workhorse for crewed and cargo missions to the , with over 2,000 historical launches and continued operations in 2025; its expendable stages ensure simplicity and a below 2% in recent decades, though booster parachutes allow limited post-flight analysis without reuse. Heavy-lift capabilities rest on the aging , operational since 1965 with 431 launches by 2023, but plagued by corrosion issues and a 7-10% failure rate in the due to flaws at Khrunichev; scheduled at least four Proton missions through 2025 before full retirement, transitioning to the modular family. , also expendable and engine-based, supports payloads from 3,800 kg (Angara 1.2 light variant) to 24,500 kg (A5 heavy) to ; after test flights including A5's April 2024 debut from Vostochny, operational Angara 1.2 launches occurred in 2022 and March 2025, carrying military satellites, though production delays and costs exceeding 100 million USD per A5 have slowed adoption.

Asian Developments

Asia hosts several major expendable launch system programs, primarily led by state agencies in , , , and , emphasizing reliable access to for national satellites, scientific missions, and regional security needs. These systems prioritize proven solid and liquid propulsion technologies, with launch cadences reflecting strategic priorities rather than commercial pressures. 's extensive family dominates in volume, supporting over 500 launches since the 1970s, including the 5B variant dedicated to heavy-lift tasks for the modules. The series remains fully expendable, with ongoing developments like the for future manned lunar missions, tested statically in September 2025, underscoring a focus on scalability over reusability. India's Indian Space Research Organisation (ISRO) operates the (PSLV), an expendable medium-lift rocket using alternating solid and liquid stages, capable of deploying up to 1,750 kg to sun-synchronous orbits; it has achieved over 50 successful missions since 1993, including multi-satellite rideshares. The (GSLV) complements it for heavier geostationary transfer orbits, lifting 2,000-2,500 kg payloads with a cryogenic upper stage, as demonstrated in 18 launches by 2025, including the NISAR Earth-observing satellite slated for July 2025 aboard GSLV-F14. Both vehicles embody cost-effective expendability, with PSLV's modular design enabling frequent, low-failure operations vital for India's and communication constellations. Japan's , in partnership with , retired the after 46 launches by 2024, a reliable expendable system that lofted up to 6,000 kg to using LE-7A engines and solid boosters. Its successor, the , introduced in 2023, maintains expendable architecture with engines for enhanced thrust and cost reduction to approximately $50 million per launch, evidenced by the successful October 25, 2025, debut of HTV-X1 cargo vehicle on H3-24L from . H3's design flexibility, including strap-on boosters, supports JAXA's scientific payloads like , prioritizing precision over recovery amid Japan's earthquake-prone geography. South Korea's (KARI) advanced from the partially foreign-assisted to the indigenous (KSLV-II), a three-stage expendable vehicle injecting 1.5 tons into 600-800 km sun-synchronous orbits using liquid and stages. Following successful 2022 orbital demonstrations, Nuri's technology transferred to in July 2025 for commercialization, with a CAS500-3 test launch planned for November 27, 2025, from . This progression highlights Korea's self-reliance drive, though expendable nature limits it to strategic rather than high-volume applications. Israel Aerospace Industries' Shavit-2, a solid-propellant expendable launcher derived from Jericho missile heritage, has orbited Ofeq reconnaissance satellites since 1988, with a capacity of 380 kg to low Earth orbit from Palmachim site; it remains operational for defense needs, as in the Ofek-19 mission. These Asian systems collectively demonstrate expendability's enduring value for sovereign, mission-specific launches where recovery infrastructure poses logistical or security challenges.

Emerging National Efforts

maintains the Shavit-2, a three-stage solid-propellant expendable launch vehicle developed by , capable of delivering approximately 350 kg to in a retrograde trajectory due to geopolitical launch constraints from . The program, derived from the ballistic missile series, achieved its first orbital launch in 1988 and remains operational, with the most recent success on March 29, 2023, deploying the Ofek-13 . Shavit-2 launches have totaled around 13, primarily for satellites, underscoring 's independent access to amid regional threats, though its southward-only launch azimuth limits payload efficiency compared to equatorial sites. Turkey, through the Turkish Space Agency (TUA) established in 2018, pursues orbital launch independence under its National Space Program (2021-2030), focusing on domestic rocket technologies via . Efforts include the Micro-Satellite Launching System (MSLS), a small solid-fueled vehicle for suborbital or low-payload missions, with ground tests advancing but no orbital flights as of October 2025. To enable launches, Turkey initiated construction of a in in December 2024, projected to cost $350 million and support both deployments and missile testing, reflecting strategic expansion beyond domestic geography. These initiatives build on developments but face challenges in scaling to reliable expendable orbital systems amid international restrictions. In the , the (TII), affiliated with government research entities, conducted the first domestic liquid test firing on October 6, 2025, marking initial steps toward sovereign propulsion for potential expendable launch vehicles. While the UAE has launched over 10 satellites via foreign providers, national efforts emphasize engine and subsystem development rather than full vehicles, with partnerships like Aspire Space exploring larger orbital capabilities targeting 15-tonne payloads by 2030, though reusability features are under consideration. These programs prioritize in a region dominated by imports, leveraging oil-funded investments but dependent on international collaboration for integration. Brazil's Veículo Lançador de Satélites () program, aimed at a four-stage solid-propellant expendable launcher for 200-300 kg payloads, stalled after multiple failures, including a 2003 pad explosion killing 21 technicians and destroying prototypes. Revived concepts like persist in planning, but as of 2025, no operational flights have occurred, with the shifting toward international partnerships for Alcântara Launch Center utilization rather than indigenous development. This reflects persistent technical and funding hurdles in Latin American national efforts.

Future Outlook and Debates

Ongoing and Planned Vehicles

The , developed by for the , entered operational service with its third successful launch on August 12, 2025, deploying the Metop-SG A1 from the . A fourth launch is scheduled for November 4, 2025, carrying the Sentinel-1D , demonstrating the vehicle's role in sustaining Europe's independent access to amid the retirement of Ariane 5. The two- or four-booster configurations enable payloads up to 21.6 tonnes to () in its heaviest variant, with production scaled for 5-7 annual flights to support commercial and institutional missions. United Launch Alliance's , an expendable successor to the and , achieved its first U.S. certification flight on August 13, 2025, with the USSF-106 mission from , deploying navigation technology satellites. ULA anticipates approximately 10 launches in 2025 as it transitions to a Vulcan-centric fleet, leveraging the engines and V upper stage for up to 27.2 tonnes to , prioritizing national security payloads where reliability outweighs reusability. Upgrades to the upper stage are planned starting late 2025 to enhance performance for deep-space missions. Japan's H3 rocket, jointly developed by JAXA and Mitsubishi Heavy Industries, marked a successful milestone on October 26, 2025, with the launch of the HTV-X1 cargo spacecraft to the International Space Station from Tanegashima Space Center, following its debut in 2023 and succeeding the retired H-2A. Designed for cost-effective expendable operations, H3 offers flexible configurations with payloads up to 6.5 tonnes to geostationary transfer orbit, emphasizing improved reliability through solid rocket boosters and LE-9 liquid engines for government and commercial satellites. China's family remains a of expendable launches, with ongoing variants like , 3, and 4 supporting frequent missions; the Long March 12 medium-lift rocket underwent rollout preparations in 2025 for initial flights, targeting enhanced attitude control for future adaptability. The super-heavy launcher, planned for crewed lunar missions, is designed for 70 tonnes to LEO in expendable mode, with suborbital tests conducted in 2025 to validate its YF-130 engines ahead of orbital debut in the late 2020s. These vehicles underpin China's high launch cadence, exceeding 60 annually across expendable configurations despite parallel reusability efforts. South Korea's (KSLV-II) continues development with its fourth launch scheduled for November 2025, building on prior successes to affirm indigenous heavy-lift capability for up to 2.6 tonnes to , focusing on national satellite deployments. India's (formerly GSLV Mk III) sustains operational flights for multi-tonne payloads, with the in early planning stages for heavier lifts by the 2040s, relying on clustered expendable architectures for lunar ambitions rather than single super-heavy designs.

Debates on Obsolescence Versus Niche Roles

The advent of reusable launch vehicles, particularly SpaceX's Falcon 9, which achieved over 300 successful landings by mid-2025, has intensified debates on whether expendable launch systems (ELVs) are becoming obsolete, as reusability has reduced launch costs from historical averages exceeding $10,000 per kilogram to payload to under $3,000 per kilogram in competitive markets. Proponents of obsolescence argue that the economic model of expendables—discarding hardware after single use—cannot compete long-term with reuse, where amortizing development costs over multiple flights yields up to 65% savings per launch, as modeled in analyses of vehicle lifecycle economics. This shift is evident in commercial manifests, where reusable options dominate high-volume satellite deployments, prompting even traditional ELV operators like Rocket Lab to pursue partial reusability for their Electron vehicle to sustain viability. However, advocates for ELVs' continued relevance emphasize niche roles where reusability introduces inefficiencies or risks incompatible with mission needs, such as dedicated small- launches requiring precise orbital insertions without the payload penalties from systems, which can reduce capacity by 10-20% due to and . For instance, small-lift ELVs like Firefly's Alpha fulfill responsive launch demands for constellations or payloads under 1,000 kg, where the low flight rates do not justify reusable infrastructure investments, and simplicity enhances reliability for time-sensitive operations. missions further underscore this, as U.S. programs like the —certified for assured access in 2024—prioritize expendability for its higher payload fraction to geosynchronous orbits and avoidance of -induced modes, ensuring amid geopolitical tensions. Critics of full reusability dominance, including analyses from experts, note that ELVs maintain advantages in scenarios demanding maximal performance without refurbishment delays, which can extend turnaround times beyond 60 days even for mature systems like boosters. European efforts with , operational since 2024, exemplify persistence in expendables for sovereign independence, where strategic control over supply chains outweighs marginal cost differences in low-volume, high-reliability contexts. While reusables excel in mass production analogs to , first-principles assessments reveal ELVs' irreplaceability for bespoke or infrequent missions, as evidenced by ongoing procurements for vehicles like NASA's , slated for use through the 2030s despite reusability alternatives. Thus, the debate hinges not on outright replacement but on complementary roles, with ELVs likely enduring in specialized domains even as reusability captures commoditized markets.

Geopolitical and Strategic Implications

Expendable launch systems (ELVs) underpin by enabling sovereign deployment of critical space assets, such as satellites and networks, which support military operations and deterrence. In the United States, the (NSSL) program mandates certified ELVs like the for high-priority missions to ensure rapid reconstitution of space capabilities in contested environments, with Phase 3 contracts awarded in 2024 emphasizing resilient access amid threats from adversaries. Similarly, Russia's and Proton ELVs have historically launched military payloads, including navigation satellites, reinforcing its strategic posture despite reducing global partnerships. The inherent dual-use nature of ELV technologies—sharing propulsion, guidance, and reentry systems with intercontinental ballistic missiles (ICBMs)—poses proliferation risks, as civilian launch programs can mask or accelerate weapons development. The Missile Technology Control Regime (MTCR), established in 1987, restricts transfers of ELV-related components capable of delivering 500 kg payloads over 300 km, aiming to curb ballistic missile spread; for instance, North Korea's Unha satellite launcher derives from Taepodong missile designs, blurring peaceful and military intents. Israel's Shavit ELV, adapted from Jericho missile technology, exemplifies how such systems enhance deterrence while adhering to non-proliferation norms through selective international cooperation. Geopolitically, ELV capabilities foster space sovereignty, mitigating risks of foreign dependency in an era of great-power competition; Europe's and emerging ELVs were prioritized post-2022 to counter reliance on Russian launches amid the conflict, which halved Roscosmos's commercial manifest. China's series, conducting over 60 launches annually by 2024, bolsters its assertive regional influence and challenges U.S. dominance, prompting export controls and alliances like to safeguard allied access. These dynamics elevate ELVs in strategic calculations, where assured launch infrastructure deters aggression by enabling denial or rapid replacement, though escalating rivalries risk destabilizing in .

References

  1. [1]
    Expendable Launch Vehicle - an overview | ScienceDirect Topics
    Expendable launch vehicle means a launch vehicle whose propulsive stages are flown only once. Launch vehicle means a vehicle built to operate in, or place a ...
  2. [2]
    Chapter 14: Launch - NASA Science
    Nov 4, 2024 · Expendable launch vehicles, ELV, are used once. The U.S. Space Transportation System, STS, or Space Shuttle, was designed as a reuseable ...
  3. [3]
    Glenn Launch Vehicle History - NASA
    Apr 10, 2008 · For almost 30 years, the Center was responsible for the management of design, building and launch of the Atlas/Centaur and Titan/Centaur booster vehicles.
  4. [4]
    Evolved Expendable Launch Vehicle (EELV)
    The Evolved Expendable Launch Vehicle (EELV) program provides the United States affordable, reliable, and assured access to space with two families of launch ...
  5. [5]
    Launch vehicle | Types & Definition - Britannica
    ... launch vehicles to date have been designed for only a single use; they are thus called expendable launch vehicles. With costs ranging from more than 10 ...
  6. [6]
    Economic Model of Reusable vs. Expendable Launch Vehicles
    This paper presents an economic model of the cost per launch and cost per pound of both expendable and reusable launch vehicles.
  7. [7]
    14 CFR § 420.19 - general. - Law.Cornell.Edu - Cornell University
    Orbital expendable launch vehicles are further classified by weight class, based on the weight of payload the launch vehicle can place in a 100-nm orbit, as ...<|separator|>
  8. [8]
    Rockets & Launch Vehicles – Introduction to Aerospace Flight ...
    The multi-stage approach enables the payload to achieve higher velocities and altitudes with less propellant than would be possible with a single-stage rocket ...
  9. [9]
    [PDF] Rockets and Launch Vehicles
    Summary of Cold-gas Rockets. Operating Principle. Uses the thermodynamic energy contained in a compressed gas and expands the gas through a nozzle, producing ...
  10. [10]
    V2 rocket: Origin, history and spaceflight legacy | Space
    Mar 29, 2022 · The V2 rocket was the world's first large-scale liquid-propellant rocket, developed between 1936 and 1942 in Nazi Germany.Designer · Contribution to spaceflight
  11. [11]
    History of Rocketry Chapter 4 | Spaceline
    The first test launch of a V-2 occurred on June 13, 1942. The rocket pitched out of control and crashed as a result of a propellant feed system failure. The ...Missing: origins | Show results with:origins
  12. [12]
    The Military Rockets that Launched the Space Age
    Aug 9, 2023 · Forced and enslaved laborers from various Nazi concentration camp systems built the V-2. ... At the time that Germany was launching V-2 ...
  13. [13]
    V-2
    It represented an enormous quantum leap in technology, financed by Nazi Germany in a huge development program that cost at least $ 2 billion in 1944 dollars.
  14. [14]
    70 Years Ago: First Redstone Launch From Cape Canaveral - NASA
    Aug 21, 2023 · Seventy years ago, on Aug. 20, 1953, a slender white single-stage missile lifted off from Pad 4 at Cape Canaveral, Florida. This first ...
  15. [15]
    Redstone Missile | National Air and Space Museum
    In 1961, the Mercury-Redstone rocket launched the first American into space, Alan B. Shepard. As a missile, the Redstone had a range of 200-250 miles and ...
  16. [16]
    V-2 Missile | National Air and Space Museum
    Nov 6, 2023 · The first successful R-1 launch took place on 17 September 1948, called by the Russians their first "national rocket," although still looking ...
  17. [17]
    R-7 family of launchers and ICBMs - RussianSpaceWeb.com
    Sep 25, 2025 · First launched in 1957, the R-7 was the biggest leap in the world's rocketry since the German A-4. Ironically, developed to be the first ...
  18. [18]
    Milestones 1953-1960. Sputnik, 1957 - Office of the Historian
    On October 4, 1957, the Soviet Union launched the earth's first artificial satellite, Sputnik-1. The successful launch came as a shock to experts and citizens ...
  19. [19]
    75 Years Ago: First Launch of a Two-Stage Rocket - NASA
    May 12, 2023 · In July 1946, the U.S. Army conceived of the idea of a two-stage liquid-fueled rocket by placing a WAC Corporal atop a V-2, and on June 20, 1947 ...
  20. [20]
    Korolev-- Sputnik - NASA
    The R-7 ICBM Carrier Made a Sputnik Launch Feasible. Some five years later, towards the end of 1953, having redesigned the R-7 rocket to carry a heavier ...
  21. [21]
    R-7
    Sputnik 8K71PS Russian intercontinental ballistic orbital launch vehicle. Relatively unmodified R-7 ICBM test vehicles used to launch first two Sputniks. T-1 ...
  22. [22]
    Explorer 1 - NASA
    The Launch​​ Explorer 1 was carried into orbit by a Jupiter-C rocket, launched from Cape Canaveral, Florida, at 10:48 p.m. (EST) on Jan. 31, 1958.
  23. [23]
    The Rocket That Launched Sputnik and Started the Space Race
    Oct 4, 2017 · The massive rocket that delivered Sputnik-1 into orbit. This 30-meter-tall, 280-ton rocket easily dwarfed its 185-pound cargo, making it visible to stargazing ...
  24. [24]
    [PDF] Origins of the Commercial Space Industry
    Between 1963 and 1982, U.S. expendable launch vehicle (ELV) manufacturers produced vehicles only under contract to the National Aeronautics and Space ...
  25. [25]
    [PDF] NSDD 94 Commercialization of Expendable Launch Vehicles
    May 16, 1983 · The National Space Policy identified the STS as the primary launch system for the U.S.. Government. The U.S. Government is in the process of ...
  26. [26]
    [PDF] Delta II, Atlas II, and Atlas III - Los Angeles Air Force Base
    It was also used in many commercial launches. Lockheed. Martin, the developer, launched the first commercial payload to use an Atlas II on 7. December 1991 ...
  27. [27]
    Khrunichev Space Center: Celebrating 20 Years of Commercial ...
    Proton K commercial operations began on April 9, 1996 with the successful launch of the Astra 1F telecommunications satellite built by Hughes for the European ...
  28. [28]
    The Growth of Global Commercial Interests
    The 1995 formation of Sea Launch— a joint venture among Boeing, the Russian and Ukrainian makers of the relatively inexpensive and highly reliable Zenit rocket, ...Missing: commercialization post
  29. [29]
    How is China Advancing its Space Launch Capabilities?
    Moreover, China's 38 launches in 2018 stand as the highest amount in a single year by any country in the 21st century. Launches are not the only means of ...Missing: 21st | Show results with:21st
  30. [30]
    Launch vehicle - Rockets, Satellites, Propellants | Britannica
    There are many different expendable launch vehicles in use around the world today. As the two countries most active in space, the United States and Russia ...Missing: century | Show results with:century
  31. [31]
    Hanwha to lead Korea's next-generation space rocket project
    May 14, 2024 · Hanwha will lead Korea's next-generation space launch vehicle project with a mission to land on the moon by 2032.Missing: expendable | Show results with:expendable<|separator|>
  32. [32]
    Rocket Lab Has a New Rival -- With a New Rocket | The Motley Fool
    Nov 11, 2024 · Firefly boasts a big book of business for its current Alpha-class rocket, which with a one-ton payload to low Earth orbit, outclasses Rocket ...
  33. [33]
    Alpha | Firefly - Next Spaceflight
    The second stage has a single Lightning engine utilizing the same fuel and cycle type. Missions. 6. Success Rate. 50.0%. Successes. 2. Failures. 2.
  34. [34]
    Reusable Rockets vs. Disposable Rockets: Market Trends and Cost ...
    Sep 27, 2025 · Cost per Launch Comparison: A Falcon 9 reusable launch costs $67 million, while a disposable ULA Atlas V launch costs $160 million. If you ...
  35. [35]
    launch companies are betting their future on reusability - SpaceNews
    Nov 11, 2024 · More launch companies are betting their future on reusability rather than expendable rockets that previously dominated.
  36. [36]
    Sanctions and Satellites: The Space Industry After the Russo ...
    Jun 10, 2022 · In the wake of the Ukrainian conflict, Russia's role in commercial space launch is likely to be diminished, to the benefit of alternative launch ...
  37. [37]
    New Russian Sanctions Include Commercial Space Launch Activities
    The State Department is amending the International Traffic in Arms Regulations (ITAR) to include Russia in the list of countries subject to a policy of denial ...
  38. [38]
    [PDF] The Changing Dynamics of Twenty-First-Century Space Power
    Feb 26, 2019 · Abstract. Many recent assessments of space power have posited a US decline and predicted a gloomy future in comparison to China and Russia.Missing: developers | Show results with:developers
  39. [39]
    multi-stage spacecraft - Atomic Rockets
    Aug 24, 2022 · If your spacecraft design needs more delta-V a glance at the Tsiolkovsky rocket equation tells you you have to increase the exhaust velocity or ...
  40. [40]
    Why do rockets have multiple stages?
    Dec 7, 2021 · The basic reason: tossing an extra stage can be far, far, more of a mass-savings than trying to make one stage that can do everything.Specific impulse and delta-v in Tsiolkovsky's rocket equationDeveloping intuition about altitude and velocity in multi stage rocketsMore results from space.stackexchange.com
  41. [41]
    Rocket Stages - What They Are And The Number Used On Orbital ...
    Rocket staging is the process through which a launch vehicle uses multiple stages or sections, each with its own propulsion system and fuel, to reach orbit. The ...
  42. [42]
    [PDF] ares i stage separation system design certification testing
    This paper surveys historical separation system tests that have been completed in order to ensure staging of other launch vehicles. Key separation system design ...
  43. [43]
    [PDF] Simulation and Analyses of Stage Separation of Multi-Stage Launch ...
    The objective of this paper is to demonstrate the application of ConSep for the stage separation of two-stage-to- orbit (TSTO) LGBB-Bimese reusable launch ...
  44. [44]
    [PDF] Simulation and Analyses of Stage Separation Two-Stage Reusable ...
    NASA has initiated the development of methodologies, techniques and tools needed for analysis and simulation of stage separation of next generation reusable ...
  45. [45]
    Preliminary Design of Expendable and Reusable Mixed-Staged ...
    Jan 22, 2025 · Orbital launch vehicles can be either expendable or partially/fully reusable and can assume various stage configurations. Finding an optimal ...
  46. [46]
    [PDF] Rocket Propulsion Fundamentals
    Propellants are the materials that are combusted by the engine to produce thrust. • Bipropellant liquid rocket systems consist of a fuel and an oxidizer. They ...
  47. [47]
    [PDF] N91-28199 - NASA Technical Reports Server (NTRS)
    PROPULSION. SYSTEMS. The propulsion systems in the current fleet of. U.S. expendable launch vehicles were designed for ballistic missiles and government space.
  48. [48]
    [PDF] Materials for Liquid Propulsion Systems
    The most commonly used hypergolic propellants are nitric acid, nitrogen tetra oxide (NTO), hydrogen peroxide, mono-methyl hydrazine (MMH), and unsymmetric di- ...
  49. [49]
    [PDF] ROCKET RUNDOWN - United Launch Alliance
    The. Atlas V consists of a common core booster powered by an RD-180 engine, the high-energy Centaur upper stage powered by an RL10 engine and either a 4-meter ( ...
  50. [50]
    Ariane 5
    The solid booster motors propellant load was increased by 2.43 metric tons and the case was welded, for a weight saving in dry mass of 1.9 metric tons. The core ...Missing: IV | Show results with:IV
  51. [51]
    Demonstrated benefits of combining solid and liquid propulsion in ...
    Athena - Demonstrated benefits of combining solid and liquid propulsion in small ELVs ... An Investigation of Propulsion-Structure Interaction in Solid Rocket ...
  52. [52]
    [PDF] vulcan launch systems user's guide
    Oct 16, 2023 · The payload separation system is designed to impart a relative velocity sufficient to avoid re-contact between the payload and the. Centaur V ...
  53. [53]
    [PDF] EELV Secondary Payload Adapter (ESPA) - DigitalCommons@USU
    For improved safety and ease of integration, ESPA will incorporate low-shock/non-pyrotechnic secondary payload separation systems. Most importantly, ESPA.Missing: ELV | Show results with:ELV
  54. [54]
    ESPAStar - Northrop Grumman
    The ESPAStar platform uses a customized EELV Secondary Payload Adapter (ESPA) ring as part of its structure and is capable of being launched aboard any launch ...
  55. [55]
    Frangible joint design and verification - NASA Lessons Learned
    Fairings are a standard component of expendable launch vehicles, and they are always jettisoned as soon as possible after a launch vehicle has achieved an ...
  56. [56]
    [PDF] PAYLOAD USER'S GUIDE - Firefly Aerospace
    Jul 2, 2025 · The fairing separation system employs a debris free, low-shock pneumatic separation system fully tested prior to each flight. The payload ...
  57. [57]
    [PDF] NASA TM X-52399 z
    The information from the computer is used to generate guidance signals to control the vehicle. CENTAUR GUIDANCE SYSTEM. The Centaur launch vehicle is guided by ...
  58. [58]
    5.0 Guidance, Navigation, and Control - NASA
    Mar 13, 2025 · The Guidance, Navigation & Control (GNC) subsystem includes the components used for position determination and the components used by the ...
  59. [59]
    [PDF] Atlas and Delta Capabilities to Launch Crew to Low Earth Orbit
    We anticipate that this system will be similar for either Atlas or Delta, and will use the recent Atlas V Fault. Tolerant Inertial Navigation Unit (FTINU) as ...
  60. [60]
    [PDF] An Accurate Guidance Algorithm for Implementation Onboard ...
    An algorithm for guiding a launch vehicle carrying a small satellite to a sun synchronous LEO is presented. Before the launch, a nominal path and the ...
  61. [61]
    A guidance algorithm for launch to equatorial orbit - ResearchGate
    Aug 8, 2025 · Purpose – The purpose of this paper is to propose a new guidance algorithm for launching a satellite using an expendable rocket from an ...
  62. [62]
    An electromechanical actuation system for an expendable launch ...
    These systems utilize a pulse population modulated converter and field-oriented control scheme to obtain independent control of both the voltage and frequency.
  63. [63]
    [PDF] an electrical thrust vector control system for the vega launcher
    Sep 25, 2009 · The aim of the TVC is to control the flight of the launcher by controlling the direction of thrust. It is a nested loop (small loop) inside the ...
  64. [64]
    STP-3: Enhanced Navigation will bolster ULA accuracy
    Dec 1, 2021 · "The Enhanced Navigation accuracy results in a 66 percent reduction in required spacecraft delta V to reach their desired orbit," said Nick ...
  65. [65]
    Space Launch Vehicle Reliability - Tech Insider
    Of the 4378 space launches conducted worldwide between 1957 and 1999, 390 launches failed (the success rate was 91.1 percent), with an associated loss or ...
  66. [66]
    What are the odds of a successful space launch? - BBC
    May 19, 2023 · That works out to around a 4% failure rate – one in every 25 launches. But Wade says 2022 was an anomaly. Last year broke records in terms of ...
  67. [67]
    Atlas V | ULA - Next Spaceflight
    Success Rate. 99.5%. Successes. 103. Failures. 0. Success Streak. 94. Partial Failures. 1. Next Up. Atlas V 551 rocket launch. November 4, 03:36. ViaSat-3 F2.
  68. [68]
    For Final Time, ULA Launches “Most Metal” Delta IV Heavy Into History
    Apr 9, 2024 · After more than six decades of operational service, 389 flights launched and an impressive 95-percent success rate, the curtain finally fell on ...
  69. [69]
    Goodbye to Ariane 5 | National Air and Space Museum
    Jul 11, 2023 · Its 117 launches over more than 20 years boasted a 96 percent success rate, making Ariane 5 one of the more reliable launchers on the market.
  70. [70]
    Proton-M | Khrunichev - Next Spaceflight
    Details and launches for the Proton-M rocket from Khrunichev. ... Success Rate. 91.3%. Successes. 103. Failures. 8. Success Streak. 3. Partial ...
  71. [71]
    Soyuz-2 - Wikipedia
    Launch statistics​​ Since 2006, Soyuz‑2 rockets have accumulated a total of 156 launches, 151 of which were successful, yielding a 97% success rate.Soyuz 2.1v · Soyuz at the Guiana Space... · RD-107
  72. [72]
    TIL that the Russian Soyuz rocket, which has been in service since ...
    Oct 2, 2024 · The Russian Soyuz rocket, which has been in service since 1963 and has a success rate of 98%, begins its ignition process by firing wooden sticks inside the ...Falcon 9 has statistically become more reliable than Soyuz (2+FG).The success rate of Russian space launches in the last four years ...More results from www.reddit.com
  73. [73]
    Modeling Launch Vehicle Reliability Growth as Defect Elimination
    Nov 14, 2023 · The historical success and failure record of launch vehicles clearly demonstrates the presence of reliability growth over successive launches.
  74. [74]
    [PDF] Mission Success of U.S. Launch Vehicle Flights from a Propulsion ...
    Engine/motor, GNC, and staging functions are relatively sensitive to failures in other subsystems. Figure 8. System interactions matrix for all failed flights.
  75. [75]
    [PDF] Guide to Probability of Failure Analysis for New Expendable Launch ...
    Conducting valid probability of failure analyses of new expendable launch vehicles (ELVs) is essential to ensuring public safety for launches that employ risk ...
  76. [76]
    ESA - Ariane 5 ECA - European Space Agency
    A version of the Ariane 5 launcher, Ariane 5 ECA, was designed to place payloads weighing up to 9.6 tonnes into GTO. With its increased capacity, Ariane 5 ECA ...
  77. [77]
    Proton-M | MLM Nauka - Space Launch Now
    Payload Capacity · Launch Cost $65000000 · Low Earth Orbit 23000.0 kg · Geostationary Transfer Orbit 6920.0 kg · Direct Geostationary 3250.0 kg · Sun-Synchronous ...
  78. [78]
    [PDF] MISSION OVERVIEW - United Launch Alliance
    The Atlas V 401 rocket is the workhorse of the Atlas V fleet, delivering ... Performance to GTO: 4,750 kg (10,470 lb). Performance to LEO-Reference: 9,800 ...Missing: capacity | Show results with:capacity
  79. [79]
    Delta IV Heavy | NROL-44 - Space Launch Now
    Payload Capacity · Launch Cost $350000000 · Low Earth Orbit 28370.0 kg · Geostationary Transfer Orbit 14210.0 kg · Direct Geostationary 14500.0 kg · Sun-Synchronous ...
  80. [80]
    Japan's H-IIA rocket: beautiful, accurate, and on-time
    The H-IIA insertion accuracy generally falls within 1-sigma, which is one-third of the permissible variation agreed with the satellite operator prior to flight.
  81. [81]
    [PDF] Launch Vehicle Production and Operations Cost Metrics
    • Atlas V 501. – (3780 kg, GTO: 27.0 deg at 35,786 km x 185 km). – (8210 ... • Scaling in the cost/mass calculation lead to a sensitivity reduction of 4 or 5.
  82. [82]
    [PDF] The Launch Systems Operations Cost Model
    The tool uses historical shuttle and military aircraft data to predict maintenance hours, turnaround time, and other metrics based on the vehicle design, choice ...<|separator|>
  83. [83]
    Rocket Launch Costs (2020-2030): How Cheap Is Space ... - PatentPC
    Sep 28, 2025 · Compared to SpaceX's offerings, the Atlas V is significantly more expensive, with costs ranging from $110 million to $160 million per launch.
  84. [84]
    ULA's Atlas V rocket to launch USSF-51 for the United States Space ...
    Jul 2, 2024 · The cost is $153.0 million. A United Launch Alliance (ULA) Atlas V rocket will launch the USSF-51 mission for the United States Space Force's ...
  85. [85]
    How Much Does It Cost to Launch a Rocket? [By Type & Size]
    Aug 16, 2023 · There is US $7.5 million for Rocket Lab's Electron smallsat launcher, US $67 million for a Falcon 9 launch (SpaceX charges about US $55 million ...
  86. [86]
    Cost breakdown of Delta IV Heavy launch
    Feb 6, 2015 · Yet, costs of payload delivered are very high (around 50,000 USD/kg for GEO, around 4,000 USD/kg for LEO). I wonder what is the main cost ...Missing: structures | Show results with:structures
  87. [87]
    What is the cost per kg or unit volume (M3) to launch something into ...
    Aug 20, 2024 · Delta IV Heavy : 24 t, $350 M, $14,600/kg; Atlas V 551 : 18 t, $150 M, $8,300/kg; Ariane 5 : 21 t, ~$150 M, ~$7,000/kg; Falcon 9 expendable ...
  88. [88]
    Cost of Delta 4 Heavy launches is down but the real price is a secret
    May 14, 2019 · To observers, the $149 million price for one mission or $467.5 million for three Delta 4 Heavy missions did not make any sense.Missing: structures | Show results with:structures
  89. [89]
    Russia's replacement for the Proton rocket costs way too much
    Jun 29, 2020 · This booster, which debuted in the 1960s, had a base price of about $65 million, which was competitive with SpaceX's Falcon 9 booster.
  90. [90]
    [PDF] a framework for assessing the reusability of hardware (reusable rocket
    Reusing hardware saves resources, time, and money. It must be recovered and recertified for flight, and companies are pursuing this option.
  91. [91]
    starship launch costs vs expendable rocket costs.
    May 24, 2021 · Even discounting economies of scale, an expendable rocket has lower production costs vs reusable rockets, simply due to the lack of complexity ...
  92. [92]
    Reusable Launch Vehicles or Expendable Launch ... - Internet Archive
    Jun 13, 2011 · Outlays for designing, developing, researching, and engineering reusable launchers are necessarily higher than those for expendable launchers ...Missing: simplicity | Show results with:simplicity
  93. [93]
    [PDF] Big Dumb Boosters: A Low-Cost Space Transportation Option? (Part ...
    Engineering analyses suggested, for example, that relatively simple pressure-fed engines would be suitable for such a booster, replacing more complicated and ...
  94. [94]
    Quantitative Assessment of Multidisciplinary Design Models for ...
    The focus of the research is on the engineering modeling aspects, with the goal of evaluating in detail the accuracy of engineering-level methods for launch ...
  95. [95]
    Expendable vs reusable propulsion systems cost sensitivity
    One of the key trade studies that must be con- sidered when studying any new space transporta-. t i o n hardware is whether to go reusable or expen-.
  96. [96]
    The Role of Reusable Rockets in Reducing the Cost of Access to ...
    Aug 24, 2023 · Reusable rockets lower costs by reusing components, potentially slashing launch costs by a factor of 10 or more, and enabling new space ...
  97. [97]
    Soyuz 2.1a | RKK Energiya - Next Spaceflight
    Success Rate. 98.2%. Successes. 83. Failures. 1. Success Streak. 63. Partial Failures. 1. Next Up. Soyuz 2.1a rocket launch. November 26, 11:26 PM. Soyuz MS-28.
  98. [98]
    Towards Reusable Launchers - A Widening Perspective
    This article presents some possible reusable launcher design options in the context of a scenario that might lead to a technological convergence between space ...<|control11|><|separator|>
  99. [99]
    SpaceX's historic Falcon 9 success streak met a fiery end
    Jul 12, 2024 · With a total of 297 launches since Block 5's 2018 debut, the reusable craft still has a 99.7 percent success rate.
  100. [100]
    [PDF] Is It Worth It? The Economics of Reusable Space Transportation
    Numerous notable aerospace executives and experts have publicly expressed a range of opinions regarding the economic viability of reusable versus expendable ...Missing: statistics | Show results with:statistics
  101. [101]
    Space Launch Services Market Revenue to Attain USD 57.94 Bn by ...
    May 28, 2025 · The FAA found that, as of 2024, more than 60% of all orbital launches involved some reusable technology. While depending on traditional big ...
  102. [102]
    The United States Launch Market - The Journal of Space Commerce
    Aug 23, 2025 · Future fully reusable systems could achieve costs below $100 per kilogram, representing 97% cost reduction from traditional expendable systems.
  103. [103]
    Space launch market competition - Wikipedia
    In 2016, SpaceX had 30% global market share for newly awarded commercial launch contracts, in 2017 the market share reached 45%, and 65% in 2018.History · Launch contract competitive... · 2014 · Launch industry response - to...
  104. [104]
    The environmental impact of emissions from space launches
    Aug 9, 2025 · A review is presented of the environmental impacts of space launches, specifically of emissions from commonly used solid and liquid rocket propellants.
  105. [105]
    Envisioning a sustainable future for space launches - NIH
    We summarise the emission byproducts from rocket launches and discuss their involvement in chemical and radiative processes in the stratosphere, along with ...
  106. [106]
    [PDF] Impact of Spaceflight on Earth's Atmosphere: Climate, Ozone, and ...
    Rocket launches and reentering satellites and upper stages emit gases and aerosols into every layer of the atmosphere from Earth's surface to low earth orbit.
  107. [107]
    Modeling the Environmental Impact of LEO Satellite Constellations
    Rocket launches in 2019 emitted 5,820 tonnes of CO 2, 6,380 tonnes of H 2O, 280 tonnes of black carbon, 220 tonnes of nitrogen oxides, 500 tonnes of reactive ...
  108. [108]
    The Space Industry's Climate Impact: Part 2
    Dec 15, 2023 · As of 2022, NOAA estimated that rockets emitted ~1,000T of black carbon into the atmosphere each year. Comparing propellants: The emissions ...Missing: expendable | Show results with:expendable
  109. [109]
    Toward net-zero in space exploration: A review of technological and ...
    Apr 1, 2025 · Rocket launches produce CO2, black carbon, and water vapor in the atmosphere. •. Space debris threatens future missions, requiring better ...
  110. [110]
    The Climate and Ozone Impacts of Black Carbon Emissions From ...
    Jun 1, 2022 · We show that a 10 Gg/yr rocket BC emission increases stratospheric temperatures by as much as 1.5 K in the stratosphere. Changes in global circulation also ...
  111. [111]
    Impact of Rocket Launch and Space Debris Air Pollutant Emissions ...
    Jun 24, 2022 · We develop air pollutant emissions inventories for rocket launches and re‐entry of reusable components and debris in 2019 and for a speculative space tourism ...
  112. [112]
    Ozone decomposition on alumina: Implications for solid rocket motor ...
    Jul 15, 1996 · This work addresses the potential for stratospheric ozone depletion by launch vehicle solid rocket motor exhaust. Considering best estimates of ...
  113. [113]
    Near-future rocket launches could slow ozone recovery - Nature
    Jun 9, 2025 · The 2022 WMO/UNEP Scientific Assessment of Ozone Depletion noted heightened concerns about the increased frequency of civilian rocket launches ...
  114. [114]
    Projected increase in space travel may damage ozone layer
    Jun 21, 2022 · A NOAA study suggests that a significant boost in spaceflight activity may damage the protective ozone layer on the one planet where we live.Missing: peer- | Show results with:peer-
  115. [115]
    Environmental life cycle assessment of reusable launch vehicle fleets
    Specific emissions of concern such as black carbon (BC) are mostly driven by hydrocarbon fuels, engine cycles, film cooling and oxidiser to fuel (O/F) ratios, ...
  116. [116]
    Worldwide Rocket Launch Emissions 2019: An Inventory for Use in ...
    Oct 23, 2024 · The four principal propellant types in our catalog are kerosene-based, cryogenic ( ), hypergolic, and solid fuel (Table 1). We make the ...
  117. [117]
    [PDF] Position Paper Space Debris Mitigation
    Of all break-up debris currently still in orbit, 73% has been associated with rocket body fragmentations. These are generally assumed to have been caused by ...
  118. [118]
    ESA - About space debris - European Space Agency
    In more than 60 years of space activities, more than 6050 launches have resulted in some 56450 tracked objects in orbit, of which about 28160 remain in space.<|separator|>
  119. [119]
    Rocket debris poses risks to aircraft operations - Astronomy Magazine
    Mar 6, 2025 · And many upper rocket stages are abandoned in orbit and could reenter at a later date. (As of March 2025, over 2,000 rocket bodies are still in ...
  120. [120]
    [PDF] The Disposal of Spacecraft and Launch Vehicle Stages in Low Earth ...
    Spacecraft and launch vehicle stages abandoned in. Earth orbit have historically been a primary source of debris from accidental explosions.
  121. [121]
    A comprehensive assessment of rocket body related space debris ...
    The total number of filtered and evaluated objects from the standard and auxiliary catalogues amounts to 6946. These are all documented rocket bodies that have ...
  122. [122]
    [PDF] Process for Limiting Orbital Debris - NASA
    This standard helps ensure that spacecraft and launch vehicles meet acceptable standards for limiting orbital debris generation. 1.1.2. This standard is ...
  123. [123]
    [PDF] Space Debris Mitigation Guidelines of the Committee on ... - UNOOSA
    Historically, the primary sources of space debris in Earth orbits have been (a) accidental and intentional break-ups which produce long-lived debris and (b) ...
  124. [124]
    [PDF] Quarterly News - NASA Orbital Debris Program Office
    Sep 16, 2025 · For S/C and R/B, only launches and explosions from 1995 through 2022 were considered to account for passivation efforts which were a key element ...
  125. [125]
    ESA - Mitigating space debris generation
    1) Guarantee successful disposal · 2) Improve orbital clearance · 3) Avoid in-orbit collisions · 4) Avoid internal break-ups · 5) Prevent intentional release of ...
  126. [126]
    Space debris mitigation strategies and practices in geosynchronous ...
    Launches carried out with Ariane 5 result in the generation of debris that are purposely left in GTO according to mission planning. The Ariane debris mitigation ...
  127. [127]
    7 TECHNIQUES TO REDUCE THE FUTURE DEBRIS HAZARD
    Possible techniques for deorbiting or accelerating the decay of these objects include the use of retrograde propulsion, natural perturbing forces, and drag ...
  128. [128]
    Licensing and Safety Requirements for Operation of a Launch Site
    Oct 19, 2000 · These rules will provide licensed launch site operators with licensing and safety requirements to protect the public from the risks associated with activities ...
  129. [129]
    14 CFR Part 420 -- License to Operate a Launch Site - eCFR
    (1) A safe launch must possess a risk level estimated, in accordance with the requirements of this part, not to exceed an expected number of 1 × 10−4 casualties ...
  130. [130]
    14 CFR Part 417 -- Launch Safety - eCFR
    The total risk consists of risk posed by impacting inert and explosive debris, toxic release, and far field blast overpressure. The FAA will determine whether ...
  131. [131]
    14 CFR § 417.107 - Flight safety. - Law.Cornell.Edu
    (2) A launch operator may initiate flight only if the risk to any individual member of the public does not exceed a casualty expectation of 1 × 10−6 per launch ...
  132. [132]
    [PDF] Launch Site Safety Assessment Overview and Update
    Jul 20, 2016 · § 417.107 (b) Flight Safety - Public risk criteria. A launch operator may initiate the flight of a launch vehicle only if flight safety analysis ...
  133. [133]
    The 10 Rocket Launch Failures That Changed History
    Oct 25, 2022 · 10 Rocket Launch Failures That Changed History · Fatal Mismanagement · Tiny Miscalculations, Big Consequences · Rusty Nuts · Separation Anxiety.Missing: expendable public incidents
  134. [134]
    Expendable Launch Vehicle (ELV) Payload Safety Program
    Feb 24, 2014 · This program ensures coordination with the range flight safety process to address any payload-related public safety concerns. For NASA ELV ...
  135. [135]
    Defense Primer: National Security Space Launch Program
    Apr 28, 2025 · The United States has two certified launch providers for NSSL missions: Space Exploration Technologies Corporation (SpaceX, flying its ...
  136. [136]
    Atlas V - United Launch Alliance
    Both the Atlas V and the Delta IV rely on the RL10 propulsion system to power their second stages. Logging an impressive record of nearly 400 successful ...
  137. [137]
    Delta IV - United Launch Alliance
    The Delta rocket family had a remarkable success rate over six decades of flights and concluded with 389 launches. The final Delta mission signals ULA's ...
  138. [138]
    Launches > launch-nrol-70 - National Reconnaissance Office
    The Delta IV Heavy can lift 28,370 kg (62,540 lbs) to low Earth orbit and 13,810 kg (30,440 lbs) to geostationary transfer orbit. It is an all liquid-fueled ...Missing: specifications | Show results with:specifications
  139. [139]
    Space Systems Command, United Launch Alliance launch USSF ...
    Aug 13, 2025 · The USSF is leveraging the newly certified Vulcan Centaur rocket to deliver demonstrations and experiments to geosynchronous orbit on behalf of ...
  140. [140]
    Vulcan - United Launch Alliance
    Vulcan's Centaur V upper stage offers flexibility and extreme endurance enabling the most complex orbital insertions to the most challenging and exotic orbits.
  141. [141]
    Antares Rocket | Northrop Grumman
    Antares Rocket System Features · Incorporates both solid and liquid stages and flight-proven technologies · Provides up to 10,500 kg to a low-Earth orbit space ...Missing: expendable | Show results with:expendable
  142. [142]
    Minotaur Rocket - Northrop Grumman
    Minotaur I is a four-stage solid fuel space launch vehicle utilizing Minuteman rocket motors for its first and second stages.
  143. [143]
    Northrop Grumman launches first Minotaur 4 rocket from ...
    Apr 16, 2025 · Northrop Grumman launched its first flight of a Minotaur rocket from California in nearly 14 years. Onboard the four-stage Minotaur 4 rocket was the NROL-174 ...
  144. [144]
    Launch Vehicles - National Reconnaissance Office
    OSP-3 expands on OSP-2 by continuing to use excess ICBM motors, and includes potential Evolved Expendable Launch Vehicle new entrant launch vehicles.
  145. [145]
    Arianespace plans five Ariane 6 launches in 2025 ... - SpaceNews
    Feb 11, 2025 · Arianespace is working towards a Feb. 26 launch of an Ariane 6 carrying the CSO-3 reconnaissance satellite for the French military.
  146. [146]
    Ariane 6's first commercial flight a success! - Safran
    Mar 7, 2025 · On Thursday, March 6, 2025, Ariane 6 successfully carried out its maiden commercial flight from the Guiana Space Center in Kourou.
  147. [147]
    Europe's powerful Ariane 6 rocket launches for 3rd time ... - Space
    Aug 12, 2025 · Europe's Ariane 6 heavy-lift rocket launched for the third time ever tonight (Aug ... Ariane 6 rocket | Space photo of the day for Aug. 8, 2025; A ...<|separator|>
  148. [148]
    European Vega C rocket launches CO2-mapping satellite, 4 ... - Space
    Jul 25, 2025 · An Arianespace Vega C rocket launches the MicroCarb satellite and four CO3D Earth-observation satellites from Kourou, French Guiana on July 25, ...
  149. [149]
    Vega - CNES
    Apr 4, 2025 · Vega has become a stalwart for Earth-observation and imaging satellites, with a track record of 20 successful flights out of 22 (as of 2024).
  150. [150]
    How the Soyuz rocket compares with the rest – DW – 11/19/2018
    Nov 19, 2018 · As the Soyuz is an "expendable" rocket, the four engines fall back to Earth when their fuel is spent, and the main core is not reusable either.
  151. [151]
    Russia Plans at Least Four Proton Rocket Launches Until End of ...
    Apr 23, 2024 · "As of today, we plan to carry out at least four federal and commercial launches of Proton-M launch vehicles by the end of 2025," Varochko said.Missing: retirement | Show results with:retirement<|separator|>
  152. [152]
    Russia's next-generation rocket is a decade old and still flying ...
    Mar 6, 2024 · The Angara A5 can place up to 24.5 metric tons (about 54,000 pounds) into low-Earth orbit, according to Khrunichev. The expendable rocket has ...
  153. [153]
    The “Angara-1.2” launch vehicle with spacecraft has ... - YouTube
    Mar 16, 2025 · The Russian Aerospace Forces have successfully launched the light-class “Angara-1.2” carrier rocket from the Plesetsk Cosmodrome in the ...<|control11|><|separator|>
  154. [154]
    China launches new Long March-5B rocket for space station program
    Specially developed for China's manned space program, Long March-5B will be mainly used to launch the modules of the space station. The Long March-5B carrier ...
  155. [155]
    China's Long March-10 carrier rocket passes second static fire test
    Sep 12, 2025 · China successfully organized and carried out the second static fire test of the Long March-10 carrier rocket, the country's new-generation ...
  156. [156]
    China's new-generation Long March rockets to facilitate manned ...
    Dec 6, 2024 · China's new-generation manned launch vehicle, the Long March-10 carrier rocket, will enhance the country's lunar transfer orbit payload capacity.<|separator|>
  157. [157]
    PSLV - ISRO
    Sep 21, 2023 · Polar Satellite Launch Vehicle (PSLV) is the third generation launch vehicle of India. It is the first Indian launch vehicle to be equipped with liquid stages.
  158. [158]
    What is Geosynchronous Satellite Launch Vehicle (GSLV)?
    GSLV is an expendable space launch vehicle by ISRO to launch satellites into Geosynchronous Transfer Orbits. It is 49.13m tall and has a lift-off mass of 420 ...
  159. [159]
    India rolls out rocket for July 30 launch of powerful NISAR Earth ...
    Jul 29, 2025 · The NISAR launch will be the 18th liftoff to date for the GSLV, an expendable three-stage rocket that stands 169.6 feet (51.7 meters) tall. Get ...
  160. [160]
    Japan's new H3 rocket ready for another launch attempt after last ...
    Mar 5, 2023 · MHI aims to launch the H3 rocket for as low as $50 million per mission, about 50% of the cost of an H-2A rocket flight. Japan has launched 46 H- ...Missing: IIA | Show results with:IIA
  161. [161]
    H3 Launch Vehicle - Japan Aerospace Exploration Agency - JAXA
    The H3 Launch Vehicle aims at achieving three factors namely, high flexibility, high reliability, and high cost performance.Missing: expendable | Show results with:expendable
  162. [162]
  163. [163]
    ALOS-3 - Supercluster
    The H3 Launch Vehicle is an expendable launch system with two stages. It is a liquid-propellant rocket with strap-on solid rocket boosters and will be launched ...<|separator|>
  164. [164]
    Korean Launch Vehicle NURI - 한국항공우주연구원
    As a Korea three-stage launch vehicle, the Nuri can directly put a 1.5-ton application satellite into a 600-800 km solar synchronous orbit.
  165. [165]
    Hanwha Aerospace To Get Nuri KSLV-II Rocket Tech Transfer
    Jul 28, 2025 · Hanwha Aerospace has signed a technology transfer agreement with the Korea Aerospace Research Institute for the Nuri Korea Space Launch ...
  166. [166]
    S. Korea to launch Nuri space rocket on Nov. 27: KASA
    Sep 30, 2025 · The Nuri, formally the Korea Space Launch Vehicle II (KSLV-II), will be launched Nov. ... South Korea's homegrown space rocket Nuri ...
  167. [167]
    Small Satellite Launcher - Shavit - IAI
    A three-stage satellite launcher, powered by three solid-fuel rocket motors · Operational for 30 years and is used to launch the Israeli OFEQ satellites · 380Kg ...Missing: expendable | Show results with:expendable
  168. [168]
    Ofek-19 Mission (Shavit-2) - RocketLaunch.Live
    Sep 2, 2025 · Expendable, Final Launch of Vehicle, Flight-Proven Booster, Flight-Proven Spacecraft, Gravitational Science Satellite, Heliophysics, In-Flight ...
  169. [169]
    Israeli Shavit-2 successfully launches Ofek 13 military satellite
    Mar 29, 2023 · Israel has launched a Shavit-2 rocket carrying a military satellite into a retrograde orbit from its space launch base in Palmachim, on the coast of central ...
  170. [170]
    Shavit-2 Vehicle Overview - RocketLaunch.org
    Status. Active. Reusability. Not Reusable. View upcoming Shavit-2 launches. N/A. Launch Sites used by Shavit-2. Palmachim Airbase. Israel. 13 Launches.
  171. [171]
    National Space Program - Turkish Space Agency
    National Space Programme is prepared taking into account the global developments, with the aim to enable the coordinated and integrated execution of vision, ...Missing: launch | Show results with:launch
  172. [172]
    MSLS Micro-Satellite Launching System - Roketsan
    Roketsan engineers designed all of the critical technologies for the launch systems and the spacecraft, and all were produced using domestic means.
  173. [173]
    Turkey is building a spaceport in Somalia - The Economist
    Feb 6, 2025 · In December the government began work on a spaceport in Somalia, a project that has been projected to cost $350m.
  174. [174]
    Impact of Turkey's Space Program on the Security Environment in ...
    Turkey announced its National Space program for the next 10 years including sending Turkish citizens into space. The Turkish space program is the largest and ...
  175. [175]
    UAE's First Liquid Rocket Engine & Sovereign Space Capability by TII
    Oct 6, 2025 · TII successfully test-fires the UAE's first liquid rocket engine, marking a milestone in building national space and propulsion technology.
  176. [176]
    UAE 15 ton payload orbital launch vehicle | DefenceHub
    Jul 20, 2025 · The vehicle is designed to carry up to 15 tonnes to low-Earth orbit and is scheduled for its debut launch in 2030. The agreement could help the ...
  177. [177]
    UAE shaping future of Earth observation, satellites and space...
    Oct 16, 2025 · The UAE Space Agency has signed more than 30 significant agreements with major international space sector bodies, including NASA, Japan ...
  178. [178]
    Brazilian space rockets - a troubled road to space
    Brazil's VLS rocket program had three failed attempts, including a 2003 launchpad explosion that killed 21 and destroyed two satellites. The program would ...
  179. [179]
    VLS-1 | AEB - Next Spaceflight
    The VLS-1 was the Brazilian Space Agency's main satellite launch vehicle. The launch vehicle was to be capable of launching satellites into orbit.
  180. [180]
  181. [181]
    ESA - Ariane - European Space Agency
    Next Ariane 6 launches: 4 November 2025 at 18:03 local time (21:03 GMT/22:03 CET) with Sentinel-1D, flight VA265.
  182. [182]
    Next Ariane 6 launch VA264 scheduled for 12 August 2025
    Aug 12, 2025 · On August 12, 2025 at 9:37 p.m. local time (00:37 a.m. UTC, 2:37 a.m. CEST, on August 13), Arianespace will launch EUMETSAT's Metop-SGA1 ...
  183. [183]
    ULA tempers expectations for 2025 launch volume amid transition to ...
    Jul 27, 2025 · That puts ULA's new target at around 10 launches for 2025. Following the inaugural launch of ULA's new heavy lift rocket Vulcan Centaur in ...
  184. [184]
    Vulcan USSF-106 - United Launch Alliance
    Liftoff occurred from Space Launch Complex-41 at Cape Canaveral Space Force Station in Florida. Launch Date and Time: Aug. 12, 2025 at 8:56 p.m. EDT (0056 UTC).
  185. [185]
  186. [186]
    Long March 12: Rollout of China's New Medium Lift Rocket - Facebook
    Nov 26, 2024 · With better reliability in attitude control, the Long March-12 is expected to become the basic model for future reusable rockets in China. The ...<|separator|>
  187. [187]
    China's lunar rocket test marks milestone in bringing astronauts to ...
    Aug 16, 2025 · The superheavy launcher can deliver 27 tonnes to a trans-lunar orbit. Under current plans, two Long March-10 rockets will launch the crewed ...
  188. [188]
    China to debut new Long March and commercial rockets in 2025
    Jan 2, 2025 · The various launchers will be reusable, adapted for reusability or be expendable but be designed for regular, cost-effective flights. The ...
  189. [189]
    Expendable launch system - Wikipedia
    An expendable launch system (or expendable launch vehicle/ELV) is a launch vehicle that can be launched only once, after which its components are destroyedCurrent operators · China · United States · Iranian Space Agency
  190. [190]
    Next Generation Launch Vehicle - Wikipedia
    The Next Generation Launch Vehicle (NGLV) is a family of three-stage partially reusable medium to super heavy-lift launch vehicle
  191. [191]
    Reducing the Cost of Space Travel with Reusable Launch Vehicles
    Feb 12, 2024 · In addition, reusable rockets use less fuel than expendable rockets, making them comparatively better for the environment. Recent Projects and ...
  192. [192]
    [PDF] Growth of Reusable Space Technology: Commercial Opportunities ...
    Rocket Lab, initially targeted at small, expendable launch vehicles with its Electron rocket, has started to pursue partial reusability. The company has ...
  193. [193]
    Why are reusable rockets so hard to make? | World Economic Forum
    Jan 12, 2015 · Re-usability means that the Falcon 9-R carries within it greater complexity, systems and mass, reducing the overall payload and decreasing its performance.
  194. [194]
    The State of Launch 2025 - Payload Space
    Apr 2, 2025 · Small-lift vehicles like Rocket Lab's Electron and Firefly's Alpha play a crucial role in meeting specific demands that larger rockets can't ...
  195. [195]
    Why it's time to reach for full reusability - Aerospace America - AIAA
    May 1, 2021 · The large, well-established aerospace companies are still supplying expendable rockets and are developing new ones. And their cost per unit ...
  196. [196]
    The Missing Rocket: An Economic and Engineering Analysis of the ...
    For example, SpaceX's Falcon 9 can lift nearly 23 tonnes to LEO in its expendable configuration and up to 18 tonnes when partially reusable. To GEO, Falcon 9 ...
  197. [197]
    [PDF] Is it Worth It? - The Economics of Reusable Space Transportation
    Oct 20, 2016 · lower $/Kg = bigger payload? • Price paid for a 2000Kg payload is ~ $100M; Price per Kg is ~ $50,000. • Larger launch vehicle required. Program ...<|separator|>
  198. [198]
    Building Confidence in Commercial Launch for National Security ...
    May 6, 2025 · The NSSL program requires at least two families of space launch vehicles capable of delivering any national security payload, as well as a ...
  199. [199]
    [PDF] GAO-25-107228, NATIONAL SECURITY SPACE LAUNCH
    Jun 30, 2025 · According to the memorandum, doing so would optimally balance national security space priorities and maximize the use of existing capacity.
  200. [200]
    [PDF] Space, the New Geopolitical Arena: Satellites, Conflicts, and Space ...
    The emergence of space geopolitics as a central consideration in international relations highlights the critical role that outer space plays in global power ...
  201. [201]
    The Dual-Use Nature of Space Launch Vehicles and Ballistic ...
    Oct 17, 2024 · This article examines the nuances of SLV technologies developed by states, as well as the existential legal and tangible complexities of their military ...Missing: expendable | Show results with:expendable
  202. [202]
    [PDF] Missile Technology Control Regime (MTCR) and International Code ...
    Complete rocket systems including ballistic missile systems ... The MTEC reviews both munitions and dual use license applications for missile technology.
  203. [203]
    Josef Aschbacher on geopolitics and Europe's changing space debate
    Sep 16, 2025 · Josef Aschbacher outlines how the European Space Agency is positioning itself for a rapidly changing space economy.
  204. [204]
    Geopolitics and sovereignty are a driving force for space
    Apr 15, 2025 · Geopolitics and sovereignty are (once again) a driving force for the space industry · China is launching space capabilities at a rapid rate.Missing: risks | Show results with:risks
  205. [205]
    The space race is being reshaped by geopolitics, offering ...
    May 15, 2025 · The rise of protectionist policies, tariff wars, export controls and national security concerns is forcing space firms to adapt their strategies ...<|separator|>
  206. [206]
    The role of space power in geopolitical competition
    Jan 30, 2024 · This Report first offers a perspective on the strategic role and importance of space in the civil and military spheres of human activity.
  207. [207]
    Why Space Is a National Security Priority
    Feb 11, 2025 · The United States needs to restore its focus on space as a national security priority as conditions have changed and threats have multiplied.