Fact-checked by Grok 2 weeks ago

Thermal efficiency

Thermal efficiency is a dimensionless performance metric in that quantifies the effectiveness of a or in converting into useful work, defined as the ratio of the net work output to the total heat input supplied to the . It is typically expressed as a and applies to heat engines, which operate by absorbing from a high-temperature source, performing work, and rejecting waste heat to a low-temperature sink. The fundamental formula for thermal efficiency, denoted as η, is η = W / Q_H, where W is the useful work done and Q_H is the absorbed from the hot reservoir; equivalently, it can be written as η = 1 - Q_C / Q_H, with Q_C representing the rejected to the cold reservoir. In practice, thermal efficiency is constrained by the second law of thermodynamics, which prohibits 100% conversion of heat to work due to inevitable entropy increases and waste heat production. The theoretical maximum efficiency for any heat engine operating between two temperatures is given by the Carnot efficiency: η_Carnot = 1 - T_C / T_H, where T_C and T_H are the absolute temperatures (in Kelvin) of the cold and hot reservoirs, respectively. Real-world systems fall short of this limit due to irreversibilities such as friction, heat losses, and non-ideal processes, resulting in typical efficiencies of 25–35% for gasoline engines, 30–35% for diesel engines, and around 33% for nuclear power plants. Thermal efficiency is central to the analysis of various thermodynamic cycles that model practical engines. For instance, the Otto cycle, which approximates spark-ignition internal combustion engines, has an efficiency of η_Otto = 1 - (1 / r)^{γ-1}, where r is the compression ratio and γ is the specific heat ratio of the working fluid (approximately 1.4 for air). The Diesel cycle, used in compression-ignition engines, achieves slightly higher efficiency through higher compression ratios but involves constant-pressure heat addition, modifying the formula to account for a cutoff ratio. These cycles, along with others like the Rankine cycle in steam turbines, underscore efforts to optimize energy conversion in power generation, transportation, and industrial processes, where improving efficiency reduces fuel consumption and environmental impact.

Fundamentals

Definition and Principles

Thermal efficiency is a dimensionless measure of the performance of a that converts into useful work or desired , defined as the ratio of the useful output to the total input supplied to the system. This metric quantifies how effectively a device, such as a or , utilizes input while accounting for losses due to irreversibilities like or . For heat engines, thermal efficiency is commonly expressed by the formula \eta = \frac{W_\text{net}}{Q_\text{in}}, where W_\text{net} is the net work output and Q_\text{in} is the heat input from the high-temperature source. In systems focused on , such as boilers, it is calculated as the ratio of heat delivered to the to the content of the input. As a , thermal efficiency is unitless and ranges from 0 to 1, often presented as a from 0% to 100%; it cannot exceed 100% because the useful output cannot surpass the total input, in accordance with the principle. The Second Law of Thermodynamics imposes fundamental limits on achievable efficiency by prohibiting complete conversion of to work. The concept originated with Sadi Carnot's 1824 publication Reflections on the Motive Power of Fire, which analyzed the of engines and laid foundational principles for by examining the conversion of into mechanical work. A basic example is a boiler's , given by \eta = \left( \frac{\text{[heat](/page/Heat) exported to [fluid](/page/Fluid)}}{\text{[heat](/page/Heat) provided by [fuel](/page/Fuel)}} \right) \times 100\%, which typically ranges from 80% to 98.5% in modern high-efficiency systems (such as condensing boilers) depending on design and operating conditions as of 2025. For a , similar calculations assess how much of the 's successfully heats the target medium versus being lost to the environment.

Thermodynamic Foundations

The First Law of , which embodies the principle of , forms the foundational energy balance for calculating thermal efficiency in thermodynamic systems. It states that the change in of a , denoted as ΔU, equals the added to the system Q minus the work done by the system W: ΔU = Q - W (where W is the work done by the system). This equation ensures that is neither created nor destroyed, allowing efficiency to be expressed as the ratio of useful work output to heat input, such as η = W / Q_in for heat engines. The Second Law of Thermodynamics introduces fundamental limits on thermal efficiency through the concept of and the inevitability of irreversibilities. It asserts that in any energy conversion process, the total entropy of an cannot decrease and typically increases, reflecting the natural tendency toward disorder. This entropy increase arises from irreversibilities like , heat transfer across finite temperature differences, and mixing, which dissipate useful energy as , preventing complete conversion of into work. For reversible processes, the change in entropy ΔS is given by the of reversible heat transfer δQ_rev divided by temperature T: \Delta S = \int \frac{\delta Q_\text{rev}}{T} This relation highlights that even in ideal cases, balances impose constraints, as not all input can yield work without some rejection at lower temperatures. Two equivalent statements of the underscore these efficiency bounds. The Clausius statement declares that cannot spontaneously flow from a colder body to a hotter one without external work, implying that thermal processes require temperature gradients and cannot achieve perfect efficiency without auxiliary input. Complementing this, the Kelvin-Planck statement posits that no operating in a can absorb from a single and convert it entirely into work; some must always be rejected to a colder as . Together, these laws establish why real-world thermal efficiencies fall below theoretical maxima, as seen in idealized reversible like the , serving as prerequisites for analyzing limitations in all thermodynamic devices.

Heat Engines

Carnot Efficiency

The , introduced by Sadi Carnot in his 1824 work Réflexions sur la puissance motrice du feu, represents an idealized reversible for a operating between two thermal reservoirs at constant temperatures. It consists of four reversible processes: isothermal expansion of the , where heat is absorbed from the hot reservoir; adiabatic expansion, where no heat is exchanged and the fluid does further work; isothermal compression, where heat is rejected to the cold reservoir; and adiabatic compression, returning the fluid to its initial state./Thermodynamics/Thermodynamic_Cycles/Carnot_Cycle) This cycle assumes perfect reversibility, with no frictional losses or other irreversibilities, making it a theoretical for maximum efficiency./Thermodynamics/Thermodynamic_Cycles/Carnot_Cycle) The efficiency of a Carnot engine, denoted as \eta_{Carnot}, is given by the formula \eta_{Carnot} = 1 - \frac{T_c}{T_h}, where T_h is the absolute temperature of the hot reservoir and T_c is the absolute temperature of the cold reservoir, both measured in Kelvin./15:_Thermodynamics/15.04:_Carnots_Perfect_Heat_Engine-_The_Second_Law_of_Thermodynamics_Restated) This expression derives from the entropy balance in a reversible cycle, where the total change in entropy \Delta S over the complete cycle must be zero for the system to return to its initial state without net entropy production. To outline the derivation, consider the heat transfers: during isothermal expansion at T_h, the entropy change is \Delta S_h = Q_h / T_h, where Q_h > 0 is the heat absorbed; during isothermal compression at T_c, \Delta S_c = -Q_c / T_c, where Q_c > 0 is the heat rejected. The adiabatic processes contribute no entropy change due to reversibility and zero heat transfer. For the cycle, \Delta S = \Delta S_h + \Delta S_c = 0, yielding Q_h / T_h = Q_c / T_c, or Q_c / Q_h = T_c / T_h. The efficiency, defined as the net work output divided by heat input (\eta = W / Q_h = (Q_h - Q_c) / Q_h), then simplifies to \eta_{Carnot} = 1 - T_c / T_h. This result was rigorously established as the upper limit for any heat engine by William Thomson () in 1851, demonstrating that no real engine can exceed it without violating the second law of thermodynamics. The implications of Carnot efficiency are profound in : it depends solely on the reservoir temperatures, independent of the or engine design, highlighting the fundamental role of differentials in energy conversion./15:_Thermodynamics/15.04:_Carnots_Perfect_Heat_Engine-_The_Second_Law_of_Thermodynamics_Restated) Greater efficiency is achieved with a larger temperature difference (T_h - T_c), but practical constraints limit T_h by material properties and T_c by environmental conditions, underscoring why real systems operate below this ideal./15:_Thermodynamics/15.04:_Carnots_Perfect_Heat_Engine-_The_Second_Law_of_Thermodynamics_Restated)

Practical Cycle Efficiencies

Practical heat engine cycles approximate the ideal Carnot cycle but incorporate irreversible processes such as constant-volume or constant-pressure heat addition and rejection, resulting in lower thermal efficiencies. These cycles form the basis for common devices like internal combustion engines, gas turbines, and steam power plants, where efficiency depends on parameters like compression or pressure ratios. While the Carnot efficiency represents the theoretical maximum for given temperatures, practical cycles achieve 20-50% efficiency under typical operating conditions due to these simplifications. The models spark-ignition engines, such as those in gasoline automobiles, featuring isentropic compression, constant-volume addition, isentropic expansion, and constant-volume rejection. Its thermal efficiency is given by \eta_\text{Otto} = 1 - \frac{1}{r^{\gamma-1}} where r is the and \gamma is the specific heat ratio (approximately 1.4 for air). Higher s increase efficiency, but practical limits due to knocking constrain r to 8-12, yielding automotive Otto cycle efficiencies of 20-30%. The describes -ignition engines used in trucks and generators, with isentropic , constant-pressure heat addition, isentropic expansion, and constant-volume heat rejection. The efficiency formula is \eta_\text{Diesel} = 1 - \frac{1}{r^{\gamma-1}} \cdot \frac{\rho^\gamma - 1}{\gamma (\rho - 1)} where \rho is the cutoff ratio (volume ratio during heat addition). cycles allow higher ratios (14-25) without pre-ignition risk, achieving practical efficiencies of 30-40%. In the , which powers gas turbines for aircraft and power generation, the processes involve isentropic compression, constant-pressure heat addition, isentropic expansion, and constant-pressure heat rejection. The thermal efficiency is \eta_\text{Brayton} = 1 - \frac{1}{r_p^{(\gamma-1)/\gamma}} where r_p is the pressure ratio. Simple-cycle gas turbines operate at pressure ratios of 10-20, resulting in efficiencies of 30-40%, though combined cycles can exceed this by recovering exhaust heat. The underpins steam power plants, consisting of a , (constant-pressure addition), (isentropic expansion), and (constant-pressure rejection). Its is calculated as \eta_\text{Rankine} = \frac{(h_3 - h_4) - (h_2 - h_1)}{h_3 - h_2} where h denotes specific at the respective states (1: pump inlet, 2: boiler inlet, 3: turbine inlet, 4: condenser inlet). Practical Rankine cycles in or plants achieve 30-40% , limited by condenser temperatures and material constraints on boiler pressures.
CycleTypical ApplicationPractical Efficiency Range
Spark-ignition engines20-30%
Compression-ignition engines30-40%
BraytonGas turbines30-40%
RankineSteam power plants30-40%
These values reflect real-world performance, significantly below Carnot limits due to non-reversible processes in the cycles.

Sources of Inefficiency

In actual heat engines, thermal efficiency falls short of ideal cycle predictions due to a range of practical losses stemming from irreversibilities, unintended dissipation, suboptimal utilization, and inefficiencies in fluid compression and expansion. These factors collectively diminish the net work output relative to the heat input, often by substantial margins that highlight the gap between theoretical models and real-world . Irreversibilities represent a primary of loss, primarily through mechanical in components such as piston rings, bearings, and crankshafts, which dissipates as heat and reduces net work by approximately 2-11% of the input depending on engine speed and load. Throttling losses in intake valves and other flow restrictions further contribute to these irreversibilities by creating drops that waste potential work during gas exchange processes, typically accounting for 0-6% of fuel energy in representative cycles. Heat transfer losses occur via unwanted conduction, , and from high-temperature gases to cooler engine walls, , and surroundings, often amplified by inadequate materials. These losses can consume 20-22% of the 's , significantly lowering the availability of for conversion to work. In fuel-fired engines, incomplete arises from insufficient mixing of air and , resulting in unburned hydrocarbons, formation, or excess air dilution, which reduces the effective release and to 98-99.4% under typical conditions. This inefficiency directly cuts into the usable from the , with losses equating to 0.6-1.8% of total . Non-ideal behavior in pumps and compressors during intake and compression phases demands extra work input, elevating the required heat addition while deviating from isentropic ideals; such losses, tied to gas exchange inefficiencies, can add 0-6% to the energy penalty in low-load operations. Cumulatively, these sources of inefficiency can lower actual thermal efficiency by 20-50% relative to ideal cycle estimates, underscoring the challenges in practical implementation. For instance, a typical gasoline Otto engine operates at 20-25% brake thermal efficiency, far below the ~60% predicted for an ideal Otto cycle with a compression ratio of 10. Basic mitigation strategies include deploying low-friction advanced materials like coatings on pistons to curb mechanical losses and incorporating regenerative cycle elements, such as , to recapture internal without relying on external exchangers. These approaches can recover 5-10% of lost in optimized designs.

Heat Pumps and Refrigerators

Coefficient of Performance

The (COP) quantifies the effectiveness of heat pumps and refrigerators by expressing the ratio of the desired —either heating provided or cooling achieved—to the work input required to accomplish it. For heat pumps operating in heating mode, this is defined as the net heating capacity divided by the effective power input, typically measured in watts per watt (W/W). For refrigerators operating in cooling mode, COP is similarly the net divided by the power input. This metric evaluates how efficiently these devices move heat from a low-temperature source to a high-temperature sink using mechanical work, rather than generating heat or cold directly. The formulas for COP are derived from the first law of thermodynamics applied to the refrigeration cycle. In heating mode for a heat pump, COP_h = Q_h / W, where Q_h is the heat delivered to the hot space and W is the work input to the compressor. In cooling mode for a refrigerator, COP_c = Q_c / W, where Q_c is the heat extracted from the cold space. From the energy balance Q_h = Q_c + W, it follows that COP_h = Q_h / W = (Q_c + W) / W = 1 + (Q_c / W) = 1 + \text{COP}_c. Typical COP values range from 2 to 5 for heat pumps under standard conditions and 1 to 4 for refrigerators, reflecting practical limitations like temperature differences and system losses. As of 2025, advanced variable-speed heat pumps achieve COPs exceeding 5 in optimal conditions due to improvements in compressor design and refrigerants. Unlike thermal efficiency in heat engines, which measures useful work output relative to input and is always less than 1, COP can exceed 1 (or 100%) because it includes both the work input and the low-grade amplified from the environment, providing more useful thermal energy than the electrical work supplied. This distinction highlights COP's focus on amplification rather than conversion to mechanical work. In residential applications, air-source heat pumps, which draw from outdoor air, typically achieve COPs of 2.5 to 4.5, while ground-source heat pumps, using stable subsurface temperatures, attain higher values of 3.0 to 5.0, offering greater efficiency in colder climates. The seasonal coefficient of performance (SCOP) extends this metric by averaging COP over an entire heating or cooling season, accounting for variable loads, temperatures, and part-load operation to better represent real-world performance; for instance, SCOP is calculated as the annual heating demand divided by the annual . Measurement standards, such as ISO 13256 for water-source heat pumps and Standard 37 for testing procedures, emphasize steady-state ratings at fixed conditions alongside part-load evaluations to ensure comparable and reliable assessments. These devices operate on the principle of the reversed , which provides the theoretical upper limit for COP.

Reversed Cycles and Limits

The reversed represents the ideal thermodynamic model for refrigerators and heat pumps, achieved by reversing the direction of processes in the standard cycle to transfer heat from a lower- to a higher- one. It comprises four reversible processes: reversible adiabatic compression of the , raising its temperature from T_c to T_h; reversible isothermal heat rejection at T_h to the hot ; reversible adiabatic expansion, lowering the temperature back to T_c; and reversible isothermal heat absorption at T_c from the cold . This cycle assumes no internal irreversibilities, making it the theoretical benchmark for maximum performance in heat-moving devices. The theoretical limits on performance for these devices are expressed through the Carnot coefficients of performance (COP), derived from the principles of reversibility. For a focused on heating, the Carnot COP is \mathrm{COP}_{h,\mathrm{Carnot}} = \frac{T_h}{T_h - T_c}, where T_h and T_c are the absolute temperatures (in ) of the hot and cold reservoirs. For a emphasizing cooling, it is \mathrm{COP}_{c,\mathrm{Carnot}} = \frac{T_c}{T_h - T_c}. These limits arise from the entropy balance in a reversible , where the net change is zero: \Delta S = 0 = \frac{Q_h}{T_h} - \frac{Q_c}{T_c}, implying \frac{Q_h}{Q_c} = \frac{T_h}{T_c}, with Q_h and Q_c as the magnitudes of heat rejected and absorbed. Applying the first law of thermodynamics, W = Q_h - Q_c (work input), yields the COP expressions. The dependence on temperature lift (T_h - T_c) shows that larger differences reduce COP, as more work is required to overcome the thermodynamic . For the same reservoir temperatures, the heating COP of the reversed Carnot cycle connects directly to the efficiency of the forward Carnot heat engine: \mathrm{COP}_h = \frac{1}{1 - \eta_{\mathrm{Carnot}}}, where \eta_{\mathrm{Carnot}} = 1 - \frac{T_c}{T_h}. This relationship illustrates the symmetry between cycles that produce work from heat and those that consume work to move heat. Real-world heat pumps and refrigerators attain COP values typically 40-60% of these Carnot limits, primarily due to irreversibilities like fluid friction, finite-rate , and deviations from behavior during and . These gaps mirror those in practical heat engines, emphasizing the universal impact of non-ideal processes on thermodynamic performance.

Energy Conversion

Fuel Heating Value Effects

The higher heating value (HHV) of a represents the total released during complete , with combustion products cooled to 25°C and condensed to liquid form, thereby including the of . In contrast, the lower heating value (LHV) measures the released under the same conditions but assumes remains as vapor, excluding the recovery. This distinction arises because of hydrogen-containing fuels produces , and the to vaporize it (approximately 2,260 kJ/kg) is not always recoverable in practical systems. In combustion-based energy conversion, thermal efficiency \eta is defined as the ratio of useful work output W to the fuel's energy input, expressed as \eta = \frac{W}{m_{\text{fuel}} \times HV}, where m_{\text{fuel}} is the fuel mass and HV is the heating value. The choice of HHV or LHV significantly impacts reported efficiencies, as LHV-based calculations yield higher values—typically 5-10% greater for fuels like natural gas—since they omit the unrecovered latent heat, providing a more realistic measure for systems without exhaust condensation. This difference can lead to inconsistencies in performance comparisons across technologies unless the basis is specified. Standards organizations like ASME and ISO guide the reporting conventions to ensure comparability. ASME PTC 4 for fired steam generators (boilers) mandates HHV as the basis, reflecting the potential for full heat recovery in systems. Conversely, ASME PTC 22 for gas turbines and ISO 3977-5 prefer LHV, as these cycles rarely condense exhaust water, making HHV overly conservative. For example, coal-fired power plants conventionally report efficiencies on an LHV basis, achieving 35–45% in modern supercritical units as of due to the fuel's variable and content. directly reduces effective by absorbing for , lowering HHV by up to 1-2% per 1% increase in content, which is particularly pronounced in high- coals. A post-2000 trend in renewable energy systems has favored LHV reporting for biomass and biofuel blends, promoting consistency amid varying moisture levels (often 10-50%) that render HHV impractical, as seen in co-firing applications for power generation.

Heat Exchanger Applications

Heat exchangers play a crucial role in enhancing thermal efficiency by recovering waste heat from exhaust streams and using it to preheat incoming fluids, thereby reducing the energy required for heating and minimizing losses in thermodynamic systems. This recovery process can increase overall efficiency by 10-20% in various applications, such as power generation and industrial processes. Heat exchangers are classified into recuperative types, which enable direct heat transfer between two separate fluid streams across a dividing wall, and regenerative types, which store heat in a matrix during one phase and release it during another to preheat the incoming fluid, offering higher efficiency in cyclic operations like furnaces. Common types include shell-and-tube, plate, and finned-tube heat exchangers, each suited to specific and operating conditions. Shell-and-tube designs within a cylindrical shell for handling high-pressure fluids, while plate heat exchangers use stacked plates for compact, high-surface-area transfer ideal for viscous fluids. Finned-tube variants extend surface area with fins to improve air-side transfer in gas-liquid applications. The (ε) of a heat exchanger is defined as the ratio of actual to the maximum possible , ε = Q_actual / Q_max, where Q_max is limited by the fluid with the smaller rate; values typically range from 0.5 to 0.9 depending on and flow arrangement. For design and sizing, the number of transfer units (NTU) method is widely used, where NTU = / C_min, with U as the overall , A as the surface area, and C_min as the minimum rate of the fluids; higher NTU values indicate greater potential for . Complementing this, the log-mean temperature difference (LMTD) approach calculates heat duty as Q = × LMTD, where LMTD accounts for varying temperature differences along the exchanger: \text{LMTD} = \frac{\Delta T_1 - \Delta T_2}{\ln(\Delta T_1 / \Delta T_2)} with ΔT_1 and ΔT_2 as the temperature differences at the ends. This method is particularly effective for counterflow configurations, enabling precise prediction of performance without iterative outlet temperature assumptions. In power plants, feedwater heaters serve as closed exchangers that extract from the to preheat , recovering exhaust and boosting cycle efficiency by up to 10-15%. In internal combustion engines, turbochargers utilize exhaust-driven s to compress intake air, indirectly recovering waste to increase power output and efficiency, often augmented by exchangers. HVAC systems employ economizers as air-to-air or water-based exchangers to precondition incoming fresh air with exhaust air , reducing mechanical cooling loads by 20-50% in moderate climates. A prominent example is in combined cycle power plants, where heat recovery steam generators capture gas exhaust to produce for a , achieving overall efficiencies of 60% or more—with recent records reaching 64% as of 2024—far surpassing simple cycle plants at around 40%. Fouling, the accumulation of deposits on heat transfer surfaces from , , or , reduces effectiveness by increasing and can lower rates by 20-50% over time, necessitating regular such as chemical or mechanical brushing to restore performance. Proper , including fouling factors in sizing, and monitoring via or temperature profiles, mitigate these effects and ensure sustained efficiency.

References

  1. [1]
    Thermal efficiency - Energy Education
    May 18, 2018 · Thermal efficiency is the fraction of heat that becomes useful work, calculated as η = W/QH, where W is useful work and QH is total heat input.
  2. [2]
    Thermal Efficiency | Definition, Examples & Calculation
    The thermal efficiency, ηth, represents the fraction of heat, QH, converted to work. It is a dimensionless performance measure of a heat engine that uses ...<|control11|><|separator|>
  3. [3]
    6.2: Engines and Thermal Efficiency - Physics LibreTexts
    Nov 8, 2022 · Cyclic processes provide a means to have repeatable ways to convert heat energy that comes into the gas into work energy that leaves the gas.Simple Engine · Real-World Engines · Thermal Efficiency · Otto Cycle
  4. [4]
    Thermodynamic Efficiency - an overview | ScienceDirect Topics
    Thermodynamic efficiency is defined as the fraction of heat absorbed by a heat engine that is converted into work, expressed mathematically as η = W/Q_H, ...
  5. [5]
    Boiler - Efficiency
    Boiler efficiency (%) = 100 (heat exported by the fluid (water, steam ..) / heat provided by the fuel)
  6. [6]
    June 12, 1824: Sadi Carnot Publishes Treatise on Heat Engines
    May 26, 2009 · In 1824 he published Reflections on the Motive Power of Fire, which described a theoretical “heat engine” that produced the maximum amount of work for a given ...
  7. [7]
  8. [8]
    6.3 The second law of thermodynamics: Kelvin-Planck and Clausius ...
    Clausius statement: it is impossible to construct a device that operates in a cycle and produces no effect other than the transfer of heat from a lower- ...Missing: bounds | Show results with:bounds
  9. [9]
    [PDF] Reflections on the Motive Power of Fire by Sadi Carnot - BibNum
    His approach is original because Carnot not only sought to improve steam engines but he also tried to find a general scientific theory for heat engines, just ...
  10. [10]
    [PDF] Reflections on the motive power of heat and on machines fitted to ...
    It was before this period (in 1824) tbat Sadi had published his Rljlexion. ... Finally, we will give some thoughts which reveal the religious sentiments of Sadi ...
  11. [11]
    3.5 The Internal combustion engine (Otto Cycle) - MIT
    The ideal Otto cycle efficiency is shown as a function of the compression ratio in Figure 3.11. As the compression ratio, $ r$ , increases, $ \eta_\textrm ...Missing: textbook | Show results with:textbook
  12. [12]
    What is the efficiency of different types of power plants? - EIA
    For example, if the heat rate is 10,500 Btu, the efficiency is 33%. If the heat rate is 7,500 Btu, the efficiency is 45%. in 2023, EIA changed its methodology ...
  13. [13]
    Thermal Efficiency for Diesel Cycle | Equation | nuclear-power.com
    Typical diesel engines have 30-35% thermal efficiency, but low-speed engines can exceed 50%. Diesel has the highest efficiency of practical combustion engines.
  14. [14]
    [PDF] 8.3 INTERNAL COMBUSTION ENGINES Efficiencies of internal ...
    Internal combustion engine efficiencies range from 15-22% for small gas turbines, 35-40% for large gas turbines, 25-30% for small gas engines, and 35-45% for ...
  15. [15]
    3.7 Brayton Cycle - MIT
    The Brayton cycle thermal efficiency contains the ratio of the compressor exit temperature to atmospheric temperature, so that the ratio is not based on the ...
  16. [16]
    Theory of Rankine Cycle - Equations and Calculation - Nuclear Power
    In general, the thermal efficiency, ηth, of any heat engine is defined as the ratio of the work it does, W, to the heat input at the high temperature, QH.
  17. [17]
    [PDF] Defining engine efficiency limits - Department of Energy
    Heat loss to coolant. 30%. A combination of low temperature combustion and port insulation will permit a significant reduction in the heat loss from the ...
  18. [18]
    The scope for improving the efficiency and environmental impact of ...
    SI engined vehicles on average convert only 20-25% of fuel energy to motive power mainly because of the requirement of throttling at low loads and knock [18, 19] ...
  19. [19]
    Engine Efficiency - DieselNet
    The overall brake thermal efficiency of the engine is a product of the combustion, thermodynamic, gas exchange, and mechanical efficiency.
  20. [20]
    [PDF] Engine Friction Reduction Technologies - Department of Energy
    Jun 19, 2014 · – Legacy Vehicle Lubricants: Develop advanced lubricants (base fluids and additives) to reduce parasitic friction losses by 10% and increase ...
  21. [21]
    3.4 Refrigerators and Heat Pumps - MIT
    The Carnot cycles that have been drawn are based on ideal gas behavior. For different working media, however, they will look different. We will see an example ...
  22. [22]
  23. [23]
    [PDF] Second-Law Analysis to Improve the Energy Efficiency of ...
    The second-law efficiency ε, at component level, is low for the condenser, evaporator and compressor with values of. 9.7%, 48.8% and 62.5%, respectively. The ...
  24. [24]
    High Heating Value - an overview | ScienceDirect Topics
    Higher heating value (HHV) is defined as the amount of heat released by a unit mass or volume of fuel when combusted, with the products returning to a ...
  25. [25]
    HHV vs LHV - GCV vs NCV - ADG Efficiency
    Higher Heating Value vs. Lower Heating Value · HHV = water vapour is condensed = more heat is recovered · LHV = water vapour remains as vapour – less heat is ...
  26. [26]
    What is the difference between the “higher heating value” (HHV) and ...
    Mar 5, 2019 · The numerical difference between the LHV and HHV of a fuel is roughly equivalent to the amount of latent heat of vaporization that can be practically recovered.
  27. [27]
    LHV vs HHV Efficiency - Fulton
    The High Heating Value (HHV) efficiency calculation includes this energy while the Low Heating Value (LHV) calculation does not. Typical calorific values for ...
  28. [28]
    Understanding the Differences Between Higher Heating Value and ...
    These differing calculation methods influence how energy performance is measured and compared. There are two common standards used to calculate fuel efficiency: ...
  29. [29]
    [PDF] ASME PTC 4- iNdirect Method: Stack Loss Method
    The percent energy loss is equal to the ratio of energy per pound of fuel divided by the higher heating value. Eq-11. The final boiler efficiency is the sum of ...
  30. [30]
    [PDF] High efficiency electric power generation - Carbon Sequestration - MIT
    For gas turbine (GT) cycles, LHV is usually used both in the US and Europe. As an exception, HHV is often used for IGCC plants in the US so that comparison ...
  31. [31]
    The World's Most Efficient Coal-Fired Power Plants
    May 8, 2015 · Globally, the average efficiency of coal-fired generation is 33% HHV (higher heating value) basis or 35% LHV (lower heating value) basis.
  32. [32]
    Effects of Moisture and Hydrogen Content on the Heating Value of ...
    Apr 4, 2007 · Higher heating value (HHV) of a fuel decreases with increasing of its moisture content. The LHV of a fuel increases with increasing of its ...
  33. [33]
    Many industries use combined heat and power to improve energy ...
    Jul 27, 2016 · In some cases, such as industrial furnaces, efficiency improvements resulting from waste heat recovery can improve energy efficiency 10%–50%. A ...<|separator|>
  34. [34]
    [PDF] ME 418 Lecture 9 - Heat Exchanger Analysis & Design
    Shell and tube exchanger used for liquids. Air flow. Water or ... resistance per foot of tube length for a cross-flow heat exchanger using finned tubes.
  35. [35]
    Feedwater Heater - an overview | ScienceDirect Topics
    The feedwater heater is the common design innovation used in power plants for cycle efficiency improvement. It is a heat exchanger that extracts some steam ...
  36. [36]
    Summary of Turbocharging as a Waste Heat Recovery System for a ...
    Jul 27, 2023 · It is found that exhaust gas turbocharging is the most promising waste heat recovery technology to restore high-altitude internal combustion ...Introduction · Application of Exhaust Energy... · Research Status of Variable...
  37. [37]
    What is an Economizer and How Does it Improve Efficiency?
    Mar 2, 2021 · An economizer is a heat exchanger that aims to improve energy economy. It helps systems enter "free cooling" mode and captures heat from ...
  38. [38]
    How Gas Turbine Power Plants Work - Department of Energy
    ... combined cycle plants are likely to achieve efficiencies of 60 percent or more. When waste heat is captured from these systems for heating or industrial ...
  39. [39]
    Heat exchanger fouling model and preventive maintenance ...
    The performance reduction due to fouling is mitigated by periodic cleaning of the heat exchangers. However, during cleaning, the heat exchanger is out of ...
  40. [40]
    [PDF] Heat Exchanger Fouling and Cleaning - ECI Digital Archives
    The fouling of heat exchangers not only has a negative impact on heat transfer efficiency but also may restrict the output or production capacity of the ...