Fact-checked by Grok 2 weeks ago

Chemical energy

Chemical energy is the stored in the chemical bonds between atoms and molecules in substances, which can be released or absorbed during chemical reactions such as or . This form of is a type of that arises from the arrangement of atoms and the strength of their bonds, and it plays a fundamental role in powering biological processes, activities, and everyday technologies. When chemical bonds are broken or formed, the difference between reactants and products determines whether the reaction is exothermic (releasing ) or endothermic (absorbing ), often manifesting as , , or mechanical work. In practical terms, chemical energy is harnessed from sources like fossil fuels, , and batteries, where it is converted into other usable forms such as in engines or in cells. For instance, the of in a car engine transforms chemical energy into to propel the , while in living organisms, the breakdown of glucose during provides the chemical energy needed for bodily functions. Common examples include , , , and , all of which store vast amounts of chemical energy that can be liberated through controlled reactions. The study and application of chemical energy are central to fields like and , where principles such as of thermodynamics—stating that energy is conserved as the sum of and work—govern its transformations. Measured in units like joules (J) or calories (cal), with 1 cal equaling 4.184 J, chemical energy's efficiency and conversion rates are critical for solutions, including renewable biofuels and advanced batteries. Understanding its behavior not only explains natural phenomena but also drives innovations in and conversion to address global challenges like .

Definition and Fundamentals

Definition

Chemical energy is the potential energy stored in the arrangement of atoms within molecules, particularly in the chemical bonds that hold them together. This energy becomes available when chemical reactions break or form these bonds, allowing the rearrangement of atomic structures and the release or absorption of energy. At its core, chemical energy represents a form of derived from the electromagnetic forces acting between charged particles, such as electrons and atomic nuclei, within molecules. These interactions create stable configurations that store until disrupted by a . Everyday examples illustrate this concept: the chemical in glucose molecules provides fuel for biological processes when consumed as food, while the in gasoline powers vehicles through combustion. The foundational principle governing energy changes in chemical reactions is the first law of thermodynamics, which conserves by stating that the change in the of a system, \Delta E, equals the added to the system, q, plus the work done on the system, w: \Delta E = q + w

Relation to Other Forms of Energy

Chemical energy interconverts with other forms through various processes, enabling practical applications across natural and engineered systems. In reactions, such as the burning of wood or , chemical energy stored in molecular bonds is released and converted into , heating surrounding materials or fluids. Electrochemical devices like batteries transform chemical energy directly into via oxidation-reduction reactions between electrodes and electrolytes, powering devices from portable to electric vehicles. In biological contexts, such as , chemical energy from the of (ATP) is converted into , enabling movement through the cyclic interaction of and filaments. Chemical energy differs fundamentally from , which originates from alterations in the —such as or involving protons and neutrons—releasing vastly greater amounts of energy per reaction compared to electron rearrangements in chemical processes. Similarly, energy is a macroscopic phenomenon dependent on an object's and within a , like water held behind a , whereas chemical energy operates at the within molecular structures. Within the broader energy hierarchy, chemical energy constitutes a specific type of , arising from the electrostatic interactions in chemical bonds, and it frequently serves as an intermediary in systems where it is transformed into (e.g., motion in engines) or (e.g., heat in power plants). These transformations obey the principle of , as codified in of , which asserts that the total in a remains constant during chemical reactions, with changes manifesting only as conversions between forms.

Storage in Chemical Bonds

Nature of Chemical Bonds

Chemical bonds represent the primary mechanism for storing chemical energy, arising from the interactions of electrons between atoms that result in more stable, lower-energy configurations compared to isolated atoms. In covalent bonds, atoms share pairs of electrons, achieving greater stability through this arrangement, which serves as the main site for energy storage in molecules. Ionic bonds form via electrostatic attractions between oppositely charged ions created by , lowering the system's through these attractive forces. Metallic bonds involve s surrounding a of positive ions, providing cohesive strength and through the freedom of electron movement within the structure. The stability of these bonds—and thus the stored energy—depends on the differences between bonded atoms, which influence electron distribution and . measures an atom's ability to attract in a ; significant differences (greater than about 1.7) favor ionic , while smaller differences lead to covalent sharing. In polar covalent bonds, unequal sharing due to moderate differences creates partial charges, contributing to variations in stored energy across types. At the quantum mechanical level, chemical bonds emerge from the overlap of atomic orbitals, which allows electrons to occupy lower-energy molecular orbitals and reduces the overall of the relative to separated atoms. This orbital overlap spreads the electron wavefunction across multiple nuclei, enabling lower-energy states through decreased contributions, thereby storing chemical energy in the bonded configuration. Breaking such bonds requires energy input to return electrons to higher-energy, unbonded states.

Bond Dissociation Energies

Bond dissociation energy (BDE), also known as , is defined as the standard change required to homolytically cleave a specific in a gaseous , producing two radicals, at 298 and 1 pressure. This process involves breaking the symmetrically, with each fragment retaining one from the shared pair, and BDE serves as a quantitative measure of bond strength in chemical systems. The difference in BDE values between bonds broken and bonds formed in a provides a direct estimate of the net change, where reactions that form stronger bonds (higher BDE) than those broken release exothermically. For instance, in processes, the high BDE of O=O (498 kJ/mol) compared to the bonds in hydrocarbons results in significant energy output when stronger C=O and O-H bonds are formed. Average BDE values for common bonds, derived from experimental measurements in the gas phase, are summarized below. These averages account for variations across similar bond types and are useful for approximating energetics without precise molecular details.
Bond TypeAverage BDE (kJ/mol)
H-H432
C-H413
C-C347
C=C614
C≡C839
O-H463
O=O495
N≡N941
C=O799
Several factors influence values, including (shorter bonds generally have higher BDE due to greater orbital overlap), bond multiplicity (multiple bonds like double or triple are stronger than single bonds), and the chemical environment (gas-phase BDE are typically higher than in due to solvent stabilization of radicals). Electronegativity differences between bonded atoms also play a role, as greater can strengthen bonds through electrostatic contributions. For gas-phase reactions, the enthalpy change can be approximated using the relation: \Delta H_{\text{reaction}} \approx \sum \text{BDE}_{\text{broken}} - \sum \text{BDE}_{\text{formed}} This equation assumes negligible contributions from non-bonded interactions and provides a reliable first-order prediction of reaction energetics based on bond strengths alone.

Release in Chemical Reactions

Exothermic and Endothermic Processes

Chemical reactions involving the release or absorption of energy are classified as exothermic or endothermic based on the sign of the enthalpy change (\Delta H) at constant pressure, where a negative \Delta H indicates heat release to the surroundings and a positive \Delta H signifies heat absorption from the surroundings. This convention arises because enthalpy accounts for the heat transferred under constant pressure conditions, distinguishing the direction of energy flow in the reaction. Exothermic reactions are those in which the system releases net as , resulting in \Delta H < 0, as the products possess lower enthalpy than the reactants. A representative example is the combustion of methane (\ce{CH4 + 2O2 -> CO2 + 2H2O}), which has \Delta H = -890 /mol and liberates that can surroundings or drive processes. In such reactions, the breaking and forming of chemical bonds overall favor release, converting stored into . In contrast, endothermic reactions absorb net chemical energy from the surroundings, with \Delta H > 0, as the products have higher enthalpy than the reactants. Photosynthesis exemplifies this process, where plants convert carbon dioxide and water into glucose using light energy (\ce{6CO2 + 6H2O -> C6H12O6 + 6O2}), requiring an input of approximately 2800 kJ/mol to store energy in chemical bonds. Here, the reaction absorbs heat or radiant energy, increasing the system's internal energy content. Hess's law states that the total change for a is independent of the pathway taken, depending only on the initial and final states, and can be calculated by summing the \Delta H values of intermediate steps or using standard enthalpies of formation. For instance, the \Delta H for combustion can be derived from the formation enthalpies of \ce{CO2} and \ce{H2O} minus those of \ce{CH4} and \ce{O2}, yielding the same -890 kJ/mol value regardless of the route. This principle enables prediction of energy changes in complex reactions by breaking them into measurable components. Exothermic reactions are harnessed in practical applications for production, such as in to generate and power engines or . Endothermic reactions, conversely, find use in cooling systems, like instant cold packs that absorb for therapeutic purposes, or in endothermic syntheses that store for later release. These processes highlight how the directional flow in chemical reactions underpins technologies from power generation to chemical manufacturing.

Activation Energy and Reaction Kinetics

Activation energy, denoted as E_a, is defined as the minimum energy required for reactant molecules to reach the , where bonds begin to break and new bonds form during a . This energy barrier arises from the need to reorganize molecular structures, even in exothermic reactions where the net energy release is favorable. The represents an unstable, high-energy configuration that exists momentarily before proceeding to products./Kinetics/06:_Modeling_Reaction_Kinetics/6.03:_Reaction_Profiles/6.3.02:_Basics_of_Reaction_Profiles) The relationship between activation energy and reaction rate is quantitatively described by the Arrhenius equation: k = A e^{-E_a / RT}, where k is the rate constant, A is the representing the frequency of collisions, E_a is the , R is the , and T is the absolute temperature in . This equation demonstrates that higher temperatures increase the rate by providing more molecules with sufficient energy to surmount the E_a barrier, while a larger E_a exponentially decreases the rate. Experimentally, E_a is determined by measuring rate constants at varying temperatures and plotting \ln k versus $1/T, where the slope equals -E_a / R. Energy profile diagrams illustrate the progress of a reaction along the , plotting against the extent of reaction./Kinetics/06:_Modeling_Reaction_Kinetics/6.03:_Reaction_Profiles/6.3.02:_Basics_of_Reaction_Profiles) These diagrams show reactants at an initial , a peak representing the at height E_a above the reactants, and products at a final ; the difference between reactant and product energies is the change \Delta H. For exothermic reactions, products lie below reactants (\Delta H < 0), but the activation energy hump still must be overcome./Kinetics/06:_Modeling_Reaction_Kinetics/6.03:_Reaction_Profiles/6.3.02:_Basics_of_Reaction_Profiles) According to collision theory, reaction rates depend on the frequency and effectiveness of molecular collisions between reactants. Effective collisions require sufficient kinetic energy (at least E_a) and proper molecular orientation to form the transition state. Factors such as molecular complexity can reduce the pre-exponential factor A by limiting favorable orientations, while higher concentrations or temperatures increase collision frequency, thereby accelerating rates. Catalysts accelerate reactions by providing an alternative reaction pathway with a lower activation energy, without being consumed or altering the net \Delta H of the overall process. For instance, in biological systems or heterogeneous catalysts in industrial processes stabilize transition states through intermediate binding, allowing more molecules to react at lower temperatures. This kinetic enhancement does not affect the thermodynamic equilibrium but significantly influences the speed of energy release in chemical systems.

Thermodynamic Principles

Enthalpy and Internal Energy

Internal energy, denoted as U, represents the total energy contained within a thermodynamic system, encompassing the kinetic energies of molecular motion, potential energies from intermolecular forces, and chemical potential energies associated with molecular structure and bonding. In the context of chemical energy, this includes the energy stored in that can be released or absorbed during reactions. The change in internal energy, \Delta U, for a process at constant volume is equal to the heat transferred to the system, q_v, since no work is performed when volume is fixed. The first law of thermodynamics expresses the conservation of energy in a system as \Delta U = q + w, where q is the heat added to the system and w is the work done on the system. For processes involving expansion or compression against a constant external pressure, the work term is given by w = -P \Delta V, where P is the pressure and \Delta V is the change in volume; this accounts for the work done by the system during expansion, which is negative. Enthalpy, symbolized as H, is defined as H = U + PV, where P is the pressure and V is the volume of the system; this state function is particularly useful for analyzing chemical processes at constant pressure, common in laboratory conditions. Under constant pressure, the change in enthalpy \Delta H equals the heat transferred, q_p, simplifying the assessment of energy changes in reactions where volume work occurs. The standard enthalpy of formation, \Delta H_f^\circ, quantifies the enthalpy change when one mole of a compound is formed from its constituent elements in their standard states (pure substances at 1 bar pressure and specified temperature, usually 25°C). By convention, the standard enthalpies of formation for elements in their standard states are zero, providing a reference point for calculating reaction enthalpies via Hess's law. In chemical reactions, the enthalpy change \Delta H approximates the difference between the sum of bond dissociation energies of bonds broken and bonds formed, though corrections are necessary for phase changes and non-ideal gas behavior to align with precise thermodynamic measurements. Bond dissociation energies, expressed as average enthalpies for breaking specific bonds in the gas phase, thus serve as a foundational tool for estimating the chemical energy involved in bond rearrangements.

Gibbs Free Energy and Spontaneity

The Gibbs free energy, denoted as G, is a thermodynamic potential defined as G = H - TS, where H is the , T is the absolute temperature, and S is the of the system. This quantity represents the maximum reversible work that a system can perform at constant temperature and pressure, excluding expansion work. For chemical processes, the change in Gibbs free energy, \Delta G, serves as the key criterion for spontaneity: a process is spontaneous if \Delta G < 0, at equilibrium if \Delta G = 0, and non-spontaneous if \Delta G > 0. The relationship is expressed by the equation \Delta G = \Delta H - T \Delta S, where \Delta H is the change in enthalpy and \Delta S is the change in entropy. This formulation integrates both enthalpic (energy) and entropic (disorder) contributions to determine the thermodynamic favorability of a reaction. Entropy S quantifies the degree of disorder or randomness in a system, with \Delta S > 0 indicating an increase in disorder that favors spontaneity by making the -T \Delta S term negative. In many chemical reactions, such as those producing gaseous products from solids or liquids, entropy increases because gas molecules have greater freedom of movement, enhancing overall disorder. Under standard conditions (1 bar pressure, specified temperature), the standard Gibbs free energy change \Delta G^\circ links directly to the equilibrium constant K via \Delta G^\circ = -RT \ln K, where R is the gas constant. This equation allows thermodynamic data to predict the extent to which reactants convert to products at equilibrium, with \Delta G^\circ < 0 implying K > 1 and a product-favored reaction. The spontaneity predicted by \Delta G exhibits temperature dependence, as the -T \Delta S term amplifies with rising T. For exothermic reactions (\Delta H < 0) with \Delta S < 0, such as some precipitation processes, spontaneity is favored at low temperatures where the enthalpic term dominates. Conversely, for endothermic reactions (\Delta H > 0) with \Delta S > 0, like the dissolution of certain salts, high temperatures promote spontaneity by outweighing the positive \Delta H. This temperature sensitivity explains phenomena such as the boiling of water, where \Delta G = 0 at the boiling point, balancing the endothermic vaporization with the entropy gain from liquid-to-gas transition. In practice, the Gibbs free energy criterion enables prediction of reaction directionality under constant temperature and pressure conditions, providing insight into whether a chemical process will proceed without external input, independent of rate considerations. This is particularly useful in fields like and biochemistry for assessing feasibility, such as in designing energy-efficient syntheses or understanding metabolic pathways.

Measurement Methods

Calorimetry Techniques

encompasses a range of experimental techniques designed to quantify heat changes associated with chemical reactions, providing direct measurements of chemical energy transformations. These methods rely on the principle that heat released or absorbed by a alters the of a surrounding , whose properties are precisely known. By isolating the within a controlled apparatus, calorimeters enable the determination of thermodynamic quantities such as change (ΔU) or change (ΔH), depending on the conditions of constant volume or pressure, respectively. Bomb calorimetry operates at constant volume within a sealed, high-pressure vessel known as a , typically constructed from corrosion-resistant materials like . The sample is ignited in an oxygen atmosphere, and the heat evolved is absorbed by the surrounding water bath, whose rise is measured to calculate ΔU directly, as no work is performed on the surroundings under these conditions. This technique is particularly suited for rapid, complete reactions such as combustions, with modern variants including adiabatic (where the jacket matches the calorimeter to minimize heat loss), isoperibol (constant- surroundings), and static methods to enhance accuracy. often involves combusting a standard like to determine the calorimeter's . Solution calorimetry, conducted at constant pressure, measures ΔH for reactions occurring in liquid media, such as dissolutions or neutralizations. The reactants are mixed in an insulated vessel, often a flask or specialized , and the temperature change of the is monitored using a or . For example, in acid-base reactions, the heat of neutralization reflects the of proton transfer in aqueous environments. The system's , including that of the and apparatus, is determined through electrical or known standards, allowing precise computation of the . This is versatile for studying ionic or molecular interactions in . Differential scanning calorimetry (DSC) assesses heat flow differences between a sample and an inert reference as both are subjected to a controlled program, typically linear heating or cooling. This detects endothermic or exothermic events, such as transitions or reaction enthalpies, by measuring the power required to maintain equivalent temperatures in the sample and reference pans. DSC provides quantitative data on changes per unit mass, with applications in identifying stability and reaction in materials. High-resolution variants, like modulated DSC, separate reversible and non-reversible flows for more detailed analysis. Common error sources in calorimetry include inaccuracies in the heat capacity of the apparatus, which can lead to over- or underestimation of if not properly calibrated; incomplete reactions, resulting in lower-than-expected heat yields; and heat losses to the environment via conduction, , or , necessitating corrections like Regnault's or electrical compensation. Additionally, impurities in samples or deviations from standard states require adjustments to extrapolate to ideal conditions. Systematic errors, such as those from in setups, can be minimized through precise and multiple replicates. The foundations of modern calorimetry trace back to the late 18th century, when and developed the ice calorimeter around 1780-1783. This device measured heat by quantifying the amount of ice melted by the reaction's warmth, providing early quantitative insights into thermal effects in chemical processes and studies. Their work established as a cornerstone for thermodynamic measurements.

Heat of Combustion

The standard heat of combustion, denoted as \Delta H_c^\circ, represents the enthalpy change accompanying the complete oxidation of one mole of a substance in oxygen under standard conditions (298.15 K and 1 bar pressure), with reactants and products in their standard states, typically yielding carbon dioxide gas, liquid water, and other stable oxides. This value quantifies the maximum energy released as heat during combustion, serving as a key thermodynamic property for assessing fuel potential. For instance, the combustion of \alpha-D-glucose (\ce{C6H12O6}) has \Delta H_c^\circ = -2801.5 kJ/mol, reflecting the energy from breaking its carbon-hydrogen and carbon-oxygen bonds while forming stronger carbon-oxygen bonds in \ce{CO2} and hydrogen-oxygen bonds in \ce{H2O}. Representative values for common fuels illustrate the range of energy densities. The standard heat of combustion for n-octane (\ce{C8H18}), a major component of , is -5470.3 kJ/mol, significantly higher per mole than glucose due to its greater content and lower oxygenation. These values are typically measured using bomb calorimetry under controlled conditions to ensure complete reaction to \ce{CO2} and \ce{H2O}. A distinction exists between the higher heating value (HHV) and lower heating value (LHV) of , which affects practical ratings. The HHV assumes water in the products is liquid, capturing the full including the of (approximately 44 /mol per mole of water), whereas the LHV treats water as vapor, excluding this contribution and yielding a lower output. For fuels like , the difference is about 10%, with HHV at 890 /mol and LHV at 802 /mol. Heat of combustion data is applied to rate the energy content of fuels and biomass, facilitating design of combustion systems and efficiency comparisons. For example, biomass such as wood pellets has HHVs around 18-20 MJ/kg, lower than coal's 25-30 MJ/kg, guiding selections for power generation or heating. This metric informs fuel standardization and economic assessments in energy industries. As an illustrative equation, the standard combustion of liquid n-octane is: \ce{C8H18 (l) + 12.5 O2 (g) -> 8 CO2 (g) + 9 H2O (l)} \quad \Delta H_c^\circ = -5470 \, \text{kJ/mol} This reaction exemplifies how tabulated \Delta H_c^\circ values enable predictions of total heat release in stoichiometric combustion.

Examples and Applications

Fuels and Combustion

Fossil fuels, including coal, oil, and natural gas, serve as primary sources of chemical energy through their hydrocarbon compositions, which are derived from ancient organic matter compressed over geological timescales. Coal primarily consists of carbon-rich compounds with varying hydrogen and oxygen content, while oil (crude petroleum) and natural gas are dominated by hydrocarbons such as alkanes, with natural gas mainly comprising methane (CH₄). These fuels exhibit high energy densities, enabling efficient storage and transport of chemical potential energy; for instance, hard black coal yields 23–30 MJ/kg, crude oil 42–47 MJ/kg, and natural gas 42–55 MJ/kg upon combustion. The of fuels involves rapid oxidation of these hydrocarbons in the presence of oxygen, an that releases substantial heat and primarily produces (CO₂) and as byproducts. This , occurring at high temperatures, converts the chemical energy stored in molecular bonds into , powering applications from to . CO₂ emerges as the dominant product due to the high carbon fraction (60–90%) in these fuels, contributing to atmospheric accumulation. Alternative fuels offer renewable pathways to harness chemical energy via , mitigating reliance on finite fossil resources. Biofuels, such as derived from like corn or , are produced through of plant sugars and provide a carbon-neutral cycle when sourced sustainably, as the CO₂ released during burning offsets that absorbed during plant growth. , when combusted directly, yields only and can be generated from renewable , presenting advantages in renewability and near-zero carbon emissions compared to fossil fuels. These alternatives promote and reduced environmental impact, though scalability remains challenged by production costs and infrastructure needs. Combustion efficiency in fuel systems is often compromised by incomplete reactions, particularly under oxygen-limited conditions, leading to the formation of pollutants such as and . arises from partial oxidation of carbon, signaling poor -air mixing and reduced energy yield, while forms at high temperatures from atmospheric reacting with oxygen. These emissions not only lower but also contribute to air quality degradation and health risks, necessitating advanced control technologies like catalytic converters. Globally, of chemical fuels, predominantly -based, accounts for approximately 86% of supply as of 2024 data extending into 2025 trends, underscoring their dominant role in meeting worldwide energy demands. This reliance highlights the sector's scale, with driving over 80% of energy-related activities through processes.

Batteries and

Batteries convert chemical energy into through electrochemical cells, where reactions drive the flow of electrons from the to the via an external circuit. In these cells, oxidation occurs at the , releasing electrons, while takes place at the , accepting those electrons; this separation of half-reactions prevents direct recombination and enables controlled energy release. Electrochemical cells are classified into primary and secondary batteries based on rechargeability. Primary batteries, such as the zinc-manganese dioxide alkaline cell, are non-rechargeable and designed for single-use applications like remote controls, relying on irreversible reactions for one-time energy delivery. Secondary batteries, exemplified by lithium-ion cells, are rechargeable, allowing reversal of the processes through external electrical input to restore , making them suitable for portable electronics and electric vehicles. The cell potential, which quantifies the driving force for electron flow, is calculated as E^\circ_\text{cell} = E^\circ_\text{cathode} - E^\circ_\text{anode}, where standard reduction potentials determine the overall voltage under standard conditions. This potential relates directly to the change via the equation \Delta G = -nFE_\text{cell}, where n is the number of moles of s transferred, F is Faraday's constant (approximately 96,485 C/mol), and a negative \Delta G indicates a spontaneous capable of performing electrical work. Lithium-ion batteries offer gravimetric energy densities around 250 Wh/kg, significantly lower than gasoline's approximately 12,000 Wh/kg, highlighting the efficiency trade-offs in electrochemical versus combustion-based systems despite comparable volumetric potentials in advanced designs. However, challenges persist, including from mechanisms like dead lithium formation, which reduces capacity over cycles, and safety risks such as in 2020s batteries, where uncontrolled exothermic reactions can lead to fires exceeding 900°C.

Biological Systems

In biological systems, chemical energy is primarily stored and transferred through (ATP), which serves as the universal energy currency for cellular processes. to (ADP) and inorganic phosphate (Pi) releases approximately 30.5 kJ/mol under standard conditions, providing the required for endergonic reactions such as , , and . This reaction is highly exergonic and reversible, allowing ATP to couple with unfavorable processes to drive across all domains of life. Living organisms store chemical energy in macromolecules like carbohydrates, fats, and proteins, which vary in . Carbohydrates and proteins each yield about 4 kcal/g upon oxidation, while fats provide 9 kcal/g, making the most efficient long-term energy reserve due to their higher carbon-hydrogen content and lower . These biomolecules are broken down through catabolic pathways to generate ATP, with carbohydrates serving as the primary quick-release in most cells. Photosynthesis captures to synthesize glucose from and , storing it as chemical energy in an endergonic process represented by the equation 6CO₂ + 6H₂O → C₆H₁₂O₆ + 6O₂. This light-dependent reaction occurs in chloroplasts, where absorbs photons to drive electron transport, ultimately reducing NADP⁺ and producing ATP via . The overall efficiency of converting sunlight to in is low, typically 1-2%, limited by factors such as light absorption spectra, , and energy losses in the Calvin-Benson cycle. Cellular respiration reverses this process, oxidizing glucose to release stored chemical energy as heat and ATP in an exothermic reaction: C₆H₁₂O₆ + 6O₂ → 6CO₂ + 6H₂O. This occurs in three main stages—glycolysis in the cytosol, the Krebs cycle (tricarboxylic acid cycle) in the mitochondrial matrix, and the electron transport chain (ETC) on the inner mitochondrial membrane—yielding up to 38 ATP molecules per glucose under aerobic conditions. The ETC harnesses the energy from electron transfer to pump protons, creating a gradient that drives ATP synthase. Overall, aerobic respiration achieves about 40% efficiency in converting the free energy of glucose to ATP, far surpassing the efficiency of photosynthesis.

Historical Development

Early Concepts

The concept of chemical energy emerged from early attempts to explain combustion and heat release in reactions, beginning with the phlogiston theory proposed by Georg Ernst Stahl in the late 17th and early 18th centuries. Stahl, a German chemist (1660–1734), revived an ancient idea of a combustible principle, positing that phlogiston was a subtle, fire-like substance inherent in flammable materials and released during burning, which explained the apparent loss of mass in calcined substances. This theory dominated chemical thought until the late 18th century, framing chemical reactions as the liberation of an internal fiery essence rather than energy transformation. In the 1770s, Antoine Lavoisier overturned phlogiston with his oxygen theory of combustion, demonstrating that burning involved the combination of substances with oxygen from the air, not the release of phlogiston. Lavoisier, collaborating with Pierre-Simon Laplace, quantified the heat evolved in these reactions using early calorimetric methods, such as the ice calorimeter developed around 1780, which measured heat by the quantity of ice melted. This approach linked chemical change directly to measurable thermal output, laying groundwork for viewing reactions as sources of quantifiable energy. The , prevalent in the late , further shaped early understandings by treating as an indestructible fluid called caloric that flowed between substances during chemical processes. Lavoisier himself endorsed this view around 1787, applying it to explain heat absorption or release in reactions like and before the rise of kinetic theory in the mid-19th century. Caloric was imagined to expand materials when added, accounting for thermal effects without invoking motion of particles. By the early 1800s, empirical laws began connecting atomic properties to heat capacities, as seen in the Dulong–Petit law announced in 1819 by French physicists Pierre Louis Dulong and Alexis Thérèse Petit. This rule stated that the heat capacity of solid elements is roughly constant per atomic weight, implying atoms possess equivalent abilities to store thermal energy, which supported emerging ideas of energy distribution at the atomic level. The transition to modern energy concepts culminated in the 1840s with Julius Robert von Mayer and James Prescott Joule establishing the conservation of energy, including the mechanical equivalent of heat. Mayer, in 1842, first quantified this equivalence from physiological observations, while Joule's paddle-wheel experiments confirmed that mechanical work converts to heat at a fixed ratio, unifying chemical, thermal, and mechanical energies under a conservation principle.

Modern Advancements

The advent of in the 1920s revolutionized the understanding of chemical energy by providing a theoretical framework for calculating energies through computational methods. The formulation of the in 1926 by enabled the quantum mechanical description of molecular systems, allowing predictions of levels and strengths without relying solely on experimental data. This breakthrough shifted the field from empirical observations to precise wavefunction-based simulations, facilitating the computation of surfaces essential for analyzing chemical reactions. From the 1950s to the 1970s, the compilation of comprehensive thermodynamic databases marked a significant advancement in quantifying chemical energies accurately. The JANAF Thermochemical Tables, initiated under the Joint Army-Navy-Air Force thermochemical project in the mid-1950s and expanded through the 1970s, provided standardized enthalpies of formation (ΔH_f) and other thermodynamic properties for thousands of substances, enabling reliable predictions of reaction enthalpies and spontaneity. These tables, critically evaluated by experts, became a cornerstone for engineering applications and , reducing uncertainties in energy calculations from classical . In the 1990s and beyond, advanced further with the widespread adoption of (DFT), which simulates reaction energies efficiently without full experimental validation. DFT, building on the Hohenberg-Kohn theorems of 1964 but gaining practical prominence in the 1990s through improved exchange-correlation functionals like B3LYP, approximates the many-electron problem using , yielding accurate bond dissociation and activation energies for complex molecules. This method has transformed the study of chemical energy landscapes, allowing of catalysts and materials for . The 2020s have seen innovations in sustainable chemistry, particularly systems designed to harness for , addressing global climate challenges. Recent developments, such as carbon nitride-based photocatalysts that split into and oxygen under visible , mimic natural while using earth-abundant materials. These systems convert chemical energy from into storable fuels, reducing reliance on fossil-derived and mitigating CO2 emissions. Post-2020, the integration of artificial intelligence with chemical energy research has accelerated predictions of bond dissociation energies (BDE) and reaction outcomes through machine learning models. Graph neural networks and generative AI frameworks, trained on vast datasets of quantum calculations, predict BDEs for organic bonds with mean absolute errors below 1 kcal/mol, enabling rapid design of stable molecules for batteries and fuels. For instance, tools like the BDE Estimator from NREL forecast homolytic cleavage energies for C-H and O-O bonds in biofuels, while FlowER, a generative model from MIT, simulates reaction pathways with high fidelity, optimizing energy yields in synthetic routes.

References

  1. [1]
    Energy - FSU Chemistry & Biochemistry
    Energy is defined as the capacity to do work and and in chemistry we define Energy as the sum of both the work (w) done and the heat (q) generated or lost.
  2. [2]
    Forms of energy - U.S. Energy Information Administration (EIA)
    Chemical energy is energy stored in the bonds of atoms and molecules. Batteries, biomass, petroleum, natural gas, and coal are examples of chemical energy.
  3. [3]
    Chemical Energy - an overview | ScienceDirect Topics
    Chemical energy is a type of energy that is readily available from chemical reactions, chemical substance transformations, or biochemical processes.
  4. [4]
    How Cells Obtain Energy from Food - NCBI - NIH
    Glucose and other food molecules are broken down by controlled stepwise oxidation to provide chemical energy in the form of ATP and NADH.
  5. [5]
    Introduction to the First Law - Chemistry 301
    Therefore, we equate any change in the internal energy of a system with the sum of the heat and the work. \[\Delta U = q + w\]. Where the change in the internal ...
  6. [6]
    DOE Explains...Batteries - Department of Energy
    Batteries store and release electricity using chemistry, with a cathode, anode, and electrolyte. They convert electricity to chemical potential and back.<|control11|><|separator|>
  7. [7]
    The Energy of Muscle Contraction. III. Kinetic Energy During Cyclic ...
    Apr 7, 2021 · During muscle contraction, chemical energy is converted to mechanical energy when ATP is hydrolysed during cross-bridge cycling.
  8. [8]
    Chemical Versus Nuclear Reactions - Stanford
    Mar 24, 2018 · The fundamental difference between chemical reactions and nuclear reactions is which subatomic particles are rearranged in the transformation.
  9. [9]
    Lesson 2: Matter and Energy Objective
    Chemical energy, the energy stored in chemical bonds, is also considered a form of potential energy. The table below gives some more examples of kinetic and ...Missing: subset | Show results with:subset
  10. [10]
    first law of thermodynamics: conservation of energy - MIT
    The change in energy of a system is equal to the difference between the heat added to the system and the work done by the system.
  11. [11]
    When does the breaking of chemical bonds release energy?
    Jun 27, 2013 · The breaking of chemical bonds never releases energy to the external environment. Energy is only released when chemical bonds are formed.
  12. [12]
    [PDF] Atomic Structure & Chemical Bonding - Projects at Harvard
    In fact, the electronegativity difference between two bonded atoms determines the nature of the chemical bond that forms between them. If the electronegativity ...
  13. [13]
    Chemical Bonds
    ### Summary of Chemical Bonds (Covalent, Ionic, Metallic)
  14. [14]
    Electron Clouds and Bonds | Physics Van | Illinois
    Actually, quantum mechanics explains why that would promote bonding. A more spread-out wavefunction can be made of longer-wavelength, lower-energy waves.
  15. [15]
    [PDF] Bond dissociation energies in simple molecules
    The bond dissociation energy D° for a bond A— which is broken through the reaction. AB~> A+B is defined here as the standard-state enthalpy change for the ...
  16. [16]
    [PDF] Bond Dissociation Energies
    The bond dissociation energy (enthalpy change) for a bond A9B which is broken through the reaction. AB : A ⫹ B. is defined as the standard-state enthalpy ...
  17. [17]
    Prediction of organic homolytic bond dissociation enthalpies at near ...
    The cumulative difference between BDE values of all bonds broken and formed in a chemical reaction thus provides an estimate of the overall reaction enthalpy.
  18. [18]
    Common Bond Energies (D - Wired Chemist
    Hydrogen. Bond, D (kJ/mol), r (pm). H-H, 432, 74. H-B, 389, 119. H-C, 411, 109. H-Si, 318, 148. H-Ge, 288, 153. H-Sn, 251, 170. H-N, 386, 101. H-P, 322, 144.
  19. [19]
    Bond Energy: Types, Factors, Effects and Sample Questions
    Factors Affecting Bond Energy · Difference in Electronegativity of Bonded Atoms · Atomic Dimensions · The Length of the Bond.Comparative Difference... · Effect of Polarity on Bond Energy
  20. [20]
    [PDF] Enthalpy Part 1
    A negative value of an enthalpy change, ΔH < 0, indicates an exothermic reaction; a positive value, ΔH > 0, indicates an endothermic reaction. If the direction ...
  21. [21]
    5.2 Enthalpy
    Thus ΔH rxn < 0 for an exothermic reaction, and ΔH rxn > 0 for an endothermic reaction. In chemical reactions, bond breaking requires an input of energy and ...
  22. [22]
    10.4 First law of thermodynamics – Chemistry Fundamentals
    A negative value of an enthalpy change, ΔH < 0, indicates an exothermic reaction; a positive value, ΔH > 0, indicates an endothermic reaction. · Chemists use a ...
  23. [23]
    Energy
    Bond-dissociation enthalpies are always positive numbers because it takes energy to break a bond. When a table of bond energies is used to estimate the enthalpy ...
  24. [24]
    Exothermic and endothermic reactions - Student Academic Success
    Energy density can be displayed in several ways. For methane: molar enthalpy of combustion ∆H = - 890 kJ mol-1 (only enthalpy values have a negative sign) ...
  25. [25]
    Chemical Reactivity - MSU chemistry
    In the combustion of methane, for example, all six bonds in the ... exothermic reactions have a negative ΔH, and endothermic reactions have a positive ΔH.
  26. [26]
    [PDF] Background Information - GLOBE Observer
    Reactions either take in energy (endothermic) or give off energy (exothermic). Photosynthesis is an endothermic reaction, because it takes in energy from the ...
  27. [27]
    Energy Changes Accompanying Chemical Reactions
    ... photosynthesis. Photosynthesis is an endothermic reaction. The source of the energy for the formation of glucose is light (radiant energy), usually from the ...
  28. [28]
    10.5 Hess's Law – Chemistry Fundamentals - UCF Pressbooks
    Hess's Law states that if a process is the sum of steps, the total enthalpy change equals the sum of the enthalpy changes of those steps.
  29. [29]
    Hess's Law - Chemistry 301
    Hess's Law states that the overall enthalpy change is the sum of the enthalpy changes of the steps in the process.
  30. [30]
    [DOC] Endothermic and Exothermic Reactions for 8th Graders
    ... practical applications for these reactions. These reactions can be used to make hot and cold packs. Hot packs are packaged chemicals that when mixed or ...
  31. [31]
    Activation Energy - Chemical Kinetics
    The activation energy is the change in the internal energy of the system, which is not quite true. E a measures the change in the potential energy of a pair of ...
  32. [32]
    Chemical Reactivity - MSU chemistry
    The energy needed to raise the reactants to the transition state energy level is called the activation energy, ΔE‡. An example of a single-step exothermic ...<|control11|><|separator|>
  33. [33]
    The Activation Energy of Chemical Reactions
    The Arrhenius equation can be used to determine the activation energy for a reaction. We start by taking the natural logarithm of both sides of the equation.
  34. [34]
    Activation Energy
    The activation energy (E a ) - is the energy level that the reactant molecules must overcome before a reaction can occur.
  35. [35]
    18.5 Collision Theory and the Effect of Temperature on Reaction Rate
    Collision theory is based on the following postulates: The rate of a reaction is proportional to the rate of reactant collisions.
  36. [36]
    Collision Theory – Chemistry - JMU Libraries Pressbooks
    Collision theory provides a simple but effective explanation for the effect of many experimental parameters on reaction rates.
  37. [37]
    DOE Explains...Catalysts - Department of Energy
    Catalysts make this process more efficient by lowering the activation energy, which is the energy barrier that must be surmounted for a chemical reaction to ...
  38. [38]
    Effect of catalysts - Chemistry 302
    The effect of a catalyst is that it lowers the activation energy for a reaction. Generally, this happens because the catalyst changes the way the reaction ...
  39. [39]
    Energy, Enthalpy, and the First Law of Thermodynamics
    The change in the enthalpy of the system during a chemical reaction is equal to the change in its internal energy plus the change in the product of the pressure ...
  40. [40]
    Internal Energy - HyperPhysics
    Internal energy is defined as the energy associated with the random, disordered motion of molecules. It is separated in scale from the macroscopic ordered ...
  41. [41]
    First Law of Thermodynamics | Chem Lab - Truman ChemLab
    Apr 13, 2011 · The system's internal energy is a state function. This means that U is completely defined by the state variables of pressure (p), volume (V), ...<|separator|>
  42. [42]
    Enthalpy - Chemistry 301
    Therefore, if we define a new state function, H that is defined as. \[H = U +PV \]. We end up with the nice relation that the heat flow at constant pressure ...
  43. [43]
    WTT User's Guide - Enthalpy of Formation
    Enthalpy of Formation. The formation reaction is defined as formation of the specified compound from its constituent elements in their standard states.<|separator|>
  44. [44]
    [PDF] Lecture 15: Thermodynamics: Bond and Reaction Enthalpies
    The difference between bond enthalpies in products and reactants gives an estimate of the enthalpy of reaction ∆Hr°. B. ENTHALPIES OF REACTION. ∆Hr ° = Standard ...
  45. [45]
    Bond Enthalpies - Chemistry 301
    Bond enthalpy (also known as bond energy) is defined as the amount of energy required to break one mole of the stated bond.
  46. [46]
    Gibbs Free Energy
    It is defined by the Gibbs equation: DG = DH - TDS. For a spontaneous process at constant temperature and pressure, DG must be negative.
  47. [47]
  48. [48]
    Gibbs Free Energy
    Driving Forces and Gibbs Free Energy. Some reactions are spontaneous because they give off energy in the form of heat ( delta H < 0).
  49. [49]
  50. [50]
    [PDF] Lecture 16: Thermodynamics: Gibbs Free Energy and Entropy
    This reaction is spontaneous at temperatures. Entropy, S, is a measure of the of a system . ∆S = change in entropy. ∆S is a state function.
  51. [51]
    Equilibrium Constant from Delta G
    ln K equals e to the minus Delta G over R T. In this equation: R = 8.314 J mol-1 K-1 or 0.008314 kJ mol-1 K-1. T is the temperature on the Kelvin scale. Top ...
  52. [52]
  53. [53]
    13.4 Free Energy – Chemistry Fundamentals - UCF Pressbooks
    Temperature Dependence of Spontaneity ... The spontaneity of a process, as reflected in the arithmetic sign of its free energy change, is then determined by the ...
  54. [54]
    7.3: Heats of Reactions and Calorimetry - Chemistry LibreTexts
    Jul 12, 2023 · Calorimetry is the set of techniques used to measure enthalpy changes during chemical processes. It uses devices called calorimeters.
  55. [55]
    5.2 Calorimetry - Chemistry 2e | OpenStax
    Feb 14, 2019 · One technique we can use to measure the amount of heat involved in a chemical or physical process is known as calorimetry.
  56. [56]
    [PDF] Thermal Dynamics of Bomb Calorimeters - FAA Fire Safety
    The three principle methods of bomb calorimetry (adiabatic, isoperibol, and static jacket) differ only in their boundary conditions. For each of these methods, ...
  57. [57]
    [PDF] Bomb Calorimetry. Studies in Energy of Combustion of trical Units A ...
    The calorimeter consists of a closed can which con- tains a stirrer, a platinum resistance thermometer placed near the can wall, a combustion bomb with fuse ...
  58. [58]
    Solution Calorimetry - SERC (Carleton)
    Oct 22, 2008 · Solution calorimetric experiments involve the dissolution of a substance in a suitable solvent and measurement of the heat either taken up or ...
  59. [59]
    4: Differential Scanning Calorimetry (DSC) - Chemistry LibreTexts
    Jul 13, 2025 · DSC is very similar to DTA and gives much the same sort of information but DSC is more often used for quantitative measurement of energy changes.
  60. [60]
    Differential Scanning Calorimetry Techniques: Applications in ...
    DSC is a thermodynamical tool for direct assessment of the heat energy uptake, which occurs in a sample within a regulated increase or decrease in temperature.
  61. [61]
    Systematic Errors in an Isoperibol Solution Calorimeter Measured ...
    The condensation and evaporation of water are a potential source of error by this Method B, but not by Method A where the calorimeter temperature increases ...
  62. [62]
    Laplace's calorimeter | Opinion - Chemistry World
    Jul 10, 2016 · ... Lavoisier's private home laboratory. When the Scottish work was reported in France in 1780, Laplace suggested using Black's ice idea. He ...
  63. [63]
    The development of calorimetry and thermochemistry in Portugal
    Mar 30, 2010 · Lavoisier and P. S. Laplace, on their lecture Memoire sur la chaleur, presented the first ice calorimeter. They were the first scientists to ...
  64. [64]
    Prediction of Standard Combustion Enthalpy of Organic Compounds ...
    Sep 8, 2025 · The standard enthalpy of combustion is defined as the enthalpy change when one mole of a substance is completely burned in the presence of ...
  65. [65]
    α-D-Glucopyranose - the NIST WebBook
    ± 0.4 kJ/mol; Combustion at 293 K; Corresponding ΔfHºsolid = -1272.8 kJ/mol ... References. Symbols used in this document: ΔcH°solid, Enthalpy of combustion of ...
  66. [66]
    Octane - the NIST WebBook
    Octane's formula is C8H18, molecular weight is 114.2285, and its enthalpy of combustion is -5470.3 ± 1.6 kJ/mol.Enthalpy of combustion of... · References
  67. [67]
    [PDF] Benchmark the Fuel Cost of Steam Generation - eere.energy.gov
    Higher Versus Lower Heating Values. Fuel is sold based on its gross or higher heating value (HHV). ... This reduced value is known as the lower heating value (LHV) ...Missing: explanation | Show results with:explanation<|separator|>
  68. [68]
    Fossil Fuels - Energy Basics
    Nov 20, 2022 · Coal, oil, and natural gas are fossil fuels, which means that they are hydrocarbon compounds formed from dead plants and animals over millions of years ...
  69. [69]
    Fossil Fuel - an overview | ScienceDirect Topics
    Due to soaring carbon proportions, fossil fuel takes account of coal, oil, and natural gas. The carbon to hydrogen (C/H) ratios, composition varies like methane ...
  70. [70]
    Heat Values of Various Fuels - World Nuclear Association
    Sep 30, 2025 · Diesel fuel, 42-46 MJ/kg. Crude oil, 42-47 MJ/kg. Liquefied petroleum gas (LPG), 46-51 MJ/kg. Natural gas, 42-55 MJ/kg. Hard black coal (IEA ...Missing: composition | Show results with:composition
  71. [71]
    Biofuels and the environment - U.S. Energy Information ... - EIA
    Apr 13, 2022 · When burned, pure biofuels generally produce fewer emissions of particulates, sulfur dioxide, and air toxics than their fossil-fuel derived ...
  72. [72]
    Alternative Fuels and Advanced Vehicles
    Hydrogen is a potentially emissions-free alternative fuel that can be produced from renewable resources for use in fuel cell electric vehicles.
  73. [73]
    Biofuel production: exploring renewable energy solutions for a ...
    Oct 15, 2024 · Biofuels from renewable biomass and agricultural waste are considered more sustainable alternatives to fossil fuels with lower carbon emissions.<|separator|>
  74. [74]
    Incomplete Combustion - an overview | ScienceDirect Topics
    Normally, carbon monoxide is an indication of the efficiency of combustion. High carbon monoxide emissions mean poor combustion process. Due to the ...
  75. [75]
    [PDF] 1.4 Natural Gas Combustion - U.S. Environmental Protection Agency
    Improperly tuned boilers and boilers operating at off-design levels decrease combustion efficiency resulting in increased CO emissions. In some cases, the ...
  76. [76]
    Contrasting Summertime Trends in Vehicle Combustion Efficiency in ...
    Jan 8, 2025 · CO, a criteria pollutant that results from incomplete or inefficient combustion ... Emission ratios of CO/NOx (nitrogen oxides = NO + NO2) ...
  77. [77]
    Home | Statistical Review of World Energy
    Fossil fuels continue to underpin the energy system accounting for 86% of the energy mix.
  78. [78]
    Current Energy Landscape
    Energy Use by Resource ; World Energy Use, 2024 [561 Quads] (87% Fossil Fuels) Coal: 28% ; U.S. Energy Use, 2024 [94 Quads] (82% Fossil Fuels) Coal: 8% ...
  79. [79]
    [PDF] Lecture 25: Oxidation-Reduction and Electrochemical Cells
    Redox reactions are a major class of chemical reactions in which there is an exchange of electrons from one species to another. For example, 2Mg (s) + O2 (g) → ...
  80. [80]
    [PDF] Chapter 18: Electrochemistry
    A redox reaction involves a transfer of electrons from one species to another. This results in a change in oxidation number. Oxidation (reducing agent): loss of ...
  81. [81]
    Classification of Cells or Batteries
    Electrochemical batteries are classified into 4 broad categories. A primary cell or battery is one that cannot easily be recharged after one use, and are ...
  82. [82]
    Batteries and Fuel Cells – Chemistry - UH Pressbooks
    Dry cells and (most) alkaline batteries are examples of primary batteries. The second type is rechargeable and is called a secondary battery. Examples of ...
  83. [83]
    Energy density
    Diesel fuel has 10,700 Wh/L, 12,700 Wh/kg; gasoline 9,700 Wh/L, 12,200 Wh/kg; liquid hydrogen 2,600 Wh/L, 39,000 Wh/kg; and NiMH battery 280 Wh/L, 100 Wh/kg.
  84. [84]
    Early warning of thermal runaway based on state of safety for lithium ...
    Jun 10, 2025 · Dead lithium is a key factor in the degradation of battery performance and safety, which can cause the risk of short circuits and thermal ...
  85. [85]
    [PDF] Thermal runaway of lithium-ion batteries and hazards of abnormal ...
    We discuss the processes that can lead to thermal runaway in the context of lithium-ion batteries, a particularly common battery with a relatively high energy ...
  86. [86]
    ATP: Adenosine Triphosphate - OpenEd CUNY
    The calculated ∆G for the hydrolysis of one ATP mole into ADP and Pi is −7.3 kcal/mole (−30.5 kJ/mol). Since this calculation is true under standard conditions, ...
  87. [87]
    Physiology, Adenosine Triphosphate - StatPearls - NCBI Bookshelf
    ATP is commonly referred to as the "energy currency" of the cell, as it provides readily releasable energy in the bond between the second and third phosphate ...
  88. [88]
    [PDF] Nutrient Information
    The general Atwater factors of 4 kcal/g protein, 9 kcal/g fat, 4 kcal/g carbohydrate, and 7 kcal/g alcohol are applied to all foods regardless of their type ...
  89. [89]
    [PDF] Overall equation for photosynthesis
    •Overall equation for photosynthesis: 6 CO2 + 6 H2O --> C6H12O6 + 6 O2 ... Overall, photosynthesis is an endergonic process that stores energy in ...
  90. [90]
    [PDF] Biomass Energy - MIT OpenCourseWare
    May 4, 2020 · Efficiency of sunlight absorption/conversion during photosynthesis is 0.1-3%, but a fraction of it is lost in other products. Carbohydrates ...
  91. [91]
    EFFICIENCY OF ATP PRODUCTION
    An ATP efficiency of 38.3% is calculated under standard conditions (temperature is 298 Kelvin, pressure is 1 atm, pH is 7.0, and initial concentrations of ...Missing: percentage | Show results with:percentage<|control11|><|separator|>
  92. [92]
    Oxidation-Reduction Reactions
    The first step toward a theory of chemical reactions was taken by Georg Ernst Stahl in 1697 when he proposed the phlogiston theory, which was based on the ...
  93. [93]
    Elements and Atoms: Chapter 5 Fire and Earth: Lavoisier - Le Moyne
    Phlogiston was thought to be a principle of fire or combustibility which resided in substances which could be burned. Stahl had called it "the corporeal fire, ...
  94. [94]
    Direct Animal Calorimetry, the Underused Gold Standard for ...
    Following the pioneering calorimeter experiments of Lavoisier ... oxygen consumption and carbon dioxide production constitutes a total calorimetry model.
  95. [95]
    Teaching Heat: the Rise and Fall of the Caloric Theory
    The caloric theory interpretation was that caloric fluid originally in the battery was released along with the electric current and settled in the wire.
  96. [96]
    Petit & Dulong - Le Moyne
    ... law: The atoms of all simple bodies have exactly the same capacity for heat. If we recollect what has been said above respecting the kind of uncertainty ...
  97. [97]
    The Discovery of Energy Conservation: Mayer and Joule - Galileo
    In modern terms, a joule has to be equivalent to a fixed number of calories. Mayer was the first to spell out this 'mechanical equivalent of heat' and in 1842 ...Missing: history | Show results with:history
  98. [98]
    A History of Thermodynamics: The Missing Manual - PMC
    Beginning in the early 1700s there were fire-related ideas based on the putative conserved fire-plus-heat fluid phlogiston (having mass, and thus ponderable), ...
  99. [99]
    Quantum chemistry // University of Oldenburg
    May 23, 2025 · With the formulation of the Schrödinger equation in the mid-1920s ... Computational quantum chemistry now provides us with the tools to ...
  100. [100]
    Quantum Chemistry Calculations for Metabolomics - ACS Publications
    May 12, 2021 · In this review, we cover the major roadblocks currently facing metabolomics and discuss applications where quantum chemistry calculations offer a solution.
  101. [101]
    NIST-JANAF Thermochemical Tables
    NIST-JANAF Thermochemical Tables. NIST Standard Reference Database 13. Last Update to Data Content: 1998. DOI: 10.18434/T42S31. Search. or use the periodic ...NIST-JANAF Tables · NIST-JANAF Tables (PDF) · View · Formula indexMissing: history 1950s- 1970s ΔH_f
  102. [102]
    [PDF] JANAF Thermochemical Tables - Standard Reference Data
    The JANAF Thermochemical Tables are part of the NSRDS, providing data for technical areas, and are part of the NSRDS-NBS publication series.Missing: ΔH_f | Show results with:ΔH_f
  103. [103]
    Density functional theory: Its origins, rise to prominence, and future
    Aug 25, 2015 · This paper reviews the development of density-related methods back to the early years of quantum mechanics and follows the breakthrough in their application ...
  104. [104]
    Density functional theory across chemistry, physics and biology - PMC
    The past decades have seen density functional theory (DFT) evolve from a rising star in computational quantum chemistry to one of its major players.Missing: reaction onward
  105. [105]
    Artificial Photosynthesis Decoded: How Carbon Nitride Splits Water ...
    Jan 23, 2025 · In artificial photosynthesis, carbon nitride splits water into oxygen and hydrogen. This novel understanding of the mechanism is crucial for ...<|separator|>
  106. [106]
    Artificial photosynthesis learned from nature: New solar hydrogen ...
    Dec 2, 2024 · Artificial photosynthesis, which emulates this natural process of photosynthesis, uses sunlight to produce valuable resources, such as hydrogen, ...
  107. [107]
    Machine Learning‐Based Prediction of Bond Dissociation Energies ...
    Mar 22, 2025 · This study explores the application of machine learning to predict the bond dissociation energies (BDEs) of metal-trifluoromethyl compounds.Missing: post- | Show results with:post-
  108. [108]
    BDE Estimator
    The BDE Estimator is a machine-learning tool that predicts bond dissociation energies for single, noncyclic bonds in neutral organic molecules using C, H, O, ...
  109. [109]
    A new generative AI approach to predicting chemical reactions
    Sep 3, 2025 · The new FlowER generative AI system may improve the prediction of chemical reactions. The approach, developed at MIT, could provide ...<|control11|><|separator|>