Absorption
Absorption is the physical or chemical process by which one substance incorporates or assimilates another substance, energy, or wave into its bulk volume or structure, distinct from surface-limited phenomena like adsorption.[1] This uptake can occur across various scales and contexts, such as the assimilation of gases or liquids into solids or liquids, with applications in fields like environmental engineering.[2] In physics and chemistry, absorption commonly refers to the mechanism where electromagnetic radiation interacts with matter, converting its energy into internal forms like heat or chemical potential, often quantified by absorption coefficients that measure the efficiency of this energy transfer.[3] For instance, in spectroscopy, atomic or molecular absorption spectra reveal specific wavelengths absorbed due to electronic transitions, providing insights into material composition.[4] In biology and physiology, absorption denotes the transport of nutrients, drugs, or other solutes from the digestive tract or external environments into the bloodstream or cells via passive diffusion, active transport, or facilitated mechanisms, essential for homeostasis and metabolism.[5] This process is critical in organ systems like the intestines.[6] The term also has specialized meanings in fields such as engineering, mathematics, and economics, as detailed in subsequent sections.In Physics
Electromagnetic Absorption
Electromagnetic absorption refers to the process by which electromagnetic waves transfer their energy to matter, converting radiant energy into internal energy of atoms or molecules through mechanisms such as the photoelectric effect, where high-energy photons eject electrons from atoms, or electric dipole transitions, where oscillating electric fields induce vibrational or rotational excitations in molecules.[7][8] This interaction results in phenomena like heating, as absorbed photons cause atomic vibrations, or electronic excitation leading to photochemical reactions.[9] The foundational observations of this process trace back to Pierre Bouguer, who in 1729 described the exponential attenuation of light through a medium in his work Essai d'optique sur la gradation de la lumière, based on astronomical measurements of atmospheric extinction.[10] This was later extended by August Beer in 1852, who incorporated the concentration of absorbing substances into the absorption model through experimental studies on colored solutions, formalizing the relationship in his publication Einleitung in die höhere Optik.[11] The extent of electromagnetic absorption is quantified by the linear absorption coefficient \alpha, typically measured in units of cm^{-1}, which describes the fractional decrease in wave intensity per unit distance traveled through the medium due to energy conversion into non-radiative forms.[12] For instance, an \alpha value of 1 cm^{-1} indicates that the intensity halves after approximately 0.693 cm of propagation, reflecting the exponential nature of the attenuation. This coefficient arises from the probability of photon-matter interactions and varies with wavelength, material properties, and absorber density. In homogeneous media, the Beer-Lambert law formalizes this attenuation as I = I_0 e^{-\alpha l}, where I is the transmitted intensity, I_0 is the incident intensity, and l is the path length; in solutions or gases, \alpha = \epsilon c where \epsilon is the molar absorption coefficient (typically in M^{-1} cm^{-1}) and c is the concentration (in M). The law derives from the differential equation \frac{dI}{I} = -\alpha \, dx, which models the incremental loss of intensity proportional to the local intensity and absorber amount; integrating yields the exponential form, enabling precise quantification of absorber concentrations in applications like spectrophotometry.[13] Electromagnetic absorption manifests in two primary types: selective absorption, which occurs at specific wavelengths corresponding to atomic or molecular energy levels, producing narrow spectral lines such as the hydrogen Balmer series absorption at 410 nm (violet) or 656 nm (red); and broadband absorption, where materials like metals dissipate energy across a wide frequency range via free electron interactions, as seen in the near-total absorption of visible and infrared waves in thin metallic films due to conductive losses.[14][15] A key example of selective absorption is the ultraviolet blocking by ozone in Earth's stratosphere, where O_3 molecules absorb UV-B (280–315 nm) and UV-C (100–280 nm) radiation through electronic transitions, preventing harmful rays from reaching the surface while contributing to atmospheric heating.[16] In modern applications, the physical principles of electromagnetic absorption underpin solar cell efficiency, where semiconductors like silicon absorb photons with energies above the bandgap (e.g., ~1.1 eV for Si), generating electron-hole pairs that drive photocurrent, with optimal absorption balancing penetration depth and carrier collection.[17] Similarly, optical filters exploit selective absorption and interference to transmit desired wavelengths while attenuating others, as in dielectric multilayer stacks that achieve high reflectivity outside passbands through phase-coherent wave interactions.[18]Acoustic and Mechanical Absorption
Acoustic absorption is the process by which the kinetic energy of sound waves is transformed into heat within a propagating medium or upon interaction with boundaries. This dissipation occurs primarily through three mechanisms: viscous damping, which arises from frictional losses as air molecules shear against each other or solid surfaces; thermal conduction, where heat flows between compressed and expanded regions of the wave due to temperature gradients; and molecular relaxation, involving delayed responses in molecular vibrations or rotations that convert acoustic energy into internal thermal modes.[19][20] The absorption coefficient for sound propagation in media, denoted α_s and measured in nepers per meter, quantifies the rate of amplitude decay per unit distance and exhibits strong frequency dependence, often scaling with the square of frequency for viscous and thermal contributions in air at audible ranges.[21] In room acoustics, the dimensionless surface absorption coefficient α (ranging from 0 to 1) describes the fraction of incident sound power absorbed by materials, also frequency-dependent due to varying wave interactions with material microstructure.[22] Sabine's formula relates these to reverberation time T (in seconds) as T = \frac{0.161 V}{A}, where V is the room volume in cubic meters and A is the total absorption area in sabins, computed as A = \sum \alpha S over all surface areas S; this empirical relation assumes diffuse sound fields and linear incidence.[23] Sound absorption coefficients are determined via standardized laboratory methods, including reverberation room tests per ISO 10534-1, which measure equivalent absorption under random incidence by comparing decay times with and without the sample, and impedance tube measurements per ISO 10534-2, which use standing wave ratios or transfer functions to derive normal-incidence coefficients across frequencies from 50 Hz to 6.3 kHz.[24][25] For porous absorbers, absorption performance is influenced by material porosity, which governs the volume fraction available for airflow and energy dissipation, and airflow resistivity, a measure of viscous drag that optimizes absorption at values around 10,000–50,000 Pa·s/m³ for broadband efficacy in fibrous or cellular materials.[26][27] In solid materials, mechanical absorption manifests as energy loss during oscillatory deformation, primarily through hysteresis in viscoelastic substances, where the stress-strain loop encloses an area representing dissipated work, and internal friction from microstructural rearrangements or defects. This is modeled using the complex modulus E^* = E' + i E'', where E' (storage modulus) captures elastic energy storage and E'' (loss modulus) quantifies dissipation, with the ratio E''/E' defining the material's loss factor and thus its damping capacity.[28][29] Representative applications include open-cell foam absorbers, such as polyurethane wedges in recording studios, which achieve high absorption coefficients (α > 0.8 above 500 Hz) by trapping sound in tortuous pores for viscous and thermal losses.[30] In geophysics, seismic waves undergo mechanical absorption in Earth's viscoelastic layers, with anelastic attenuation strongest in the asthenosphere due to partial melting and high temperatures, reducing wave amplitudes by factors of e^{-π f t / Q} where Q is the quality factor (typically 100–500) and f the frequency.[31] The quantitative understanding of absorption in architectural contexts originated with Wallace Sabine's seminal 1900 experiments at Harvard, where he measured reverberation in lecture halls using organ pipes and seat cushions, establishing the absorption properties of common furnishings and deriving the foundational formula for room design.[32]In Chemistry
Molecular and Spectroscopic Absorption
Molecular absorption refers to the process by which molecules interact with electromagnetic radiation, leading to transitions between quantized energy levels. These transitions include electronic excitations in the ultraviolet-visible (UV-Vis) region, vibrational modes in the infrared (IR) region, and rotational changes in the microwave region. Electronic absorption typically involves promotion of electrons from ground to excited states, occurring at wavelengths around 200-800 nm, while vibrational absorption corresponds to stretching or bending of bonds at 400-4000 cm⁻¹, and rotational absorption involves tumbling of the molecule at frequencies near 10 GHz or 0.33 cm⁻¹. Quantum selection rules govern these processes; for pure rotational transitions in diatomic molecules, the change in rotational quantum number is ΔJ = ±1, ensuring only adjacent levels are accessible due to dipole moment requirements.[33][34] The intensity of molecular absorption is quantified by molar absorptivity (ε), expressed in units of L mol⁻¹ cm⁻¹, which measures the probability of a transition for a given concentration. This parameter relates to the Beer-Lambert law, A = εlc, where A is absorbance, l is path length, and c is concentration, emphasizing how absorption scales with molecular concentration in solution while highlighting chemical effects like solvent interactions that can shift ε values. Absorption spectra differ from emission spectra, where the former records energy uptake causing transitions, and the latter detects energy release; chromophores—structural units like conjugated π-systems—dictate absorption bands in UV-Vis, while auxochromes such as -OH or -NH₂ groups modify band position and intensity through electron donation or withdrawal, often causing bathochromic shifts.[35] In chemical applications, spectroscopic absorption enables identification of functional groups; for instance, the C=O stretch in carbonyl compounds appears as a strong IR band near 1700 cm⁻¹, allowing detection of ketones or aldehydes. Quantitative analysis in assays relies on these absorptions for measuring concentrations, such as in protein assays via UV absorbance at 280 nm due to aromatic residues. The quantum mechanical foundation involves the time-dependent Schrödinger equation for calculating transition probabilities, but for practical understanding, the Franck-Condon principle explains vibronic coupling in electronic spectra, where vertical transitions occur rapidly compared to nuclear motion, leading to vibrational progressions in absorption bands.[36][37][38] Recent advancements since the 1990s have utilized femtosecond spectroscopy to probe ultrafast absorption dynamics, resolving processes like solvent relaxation or intramolecular electron transfer on picosecond to femtosecond timescales, providing insights into non-equilibrium states following initial excitation.[39]Gas-Liquid Absorption Processes
Gas-liquid absorption is a unit operation in chemical engineering where a soluble gaseous solute is transferred from a gas phase into a liquid solvent, primarily driven by the difference in partial pressure of the solute across the gas-liquid interface. This process relies on the solubility of the gas in the liquid, often described by Henry's law, which states that at equilibrium, the partial pressure P of the solute in the gas phase is proportional to its mole fraction x in the liquid phase: P = H x, where H is Henry's constant, a measure of solubility that varies with temperature and the solute-solvent pair.[40] The driving force for mass transfer is the deviation from this equilibrium, with the rate of absorption proportional to the partial pressure gradient.[41] The theoretical foundation for modeling mass transfer in gas-liquid absorption is the two-film theory, proposed by Whitman in 1923, which assumes stagnant liquid and gas films at the interface where resistance to transfer occurs. In this model, the solute diffuses through the gas film to the interface and then through the liquid film into the bulk liquid, with the overall mass transfer coefficient K_G for the gas phase given by K_G = \frac{1}{\frac{1}{k_g} + \frac{H}{k_l}}, where k_g and k_l are the individual film mass transfer coefficients for the gas and liquid phases, respectively. This theory simplifies the prediction of absorption rates by combining resistances in series, though it assumes steady-state diffusion without convection within the films.[42] Gas-liquid absorption processes are classified into physical and chemical types based on the mechanism of solute capture. Physical absorption involves the physical dissolution of the gas into the solvent without chemical reaction, such as the absorption of carbon dioxide (CO₂) into water, where solubility is governed solely by partial pressure and temperature. In contrast, chemical absorption entails reversible chemical reactions between the solute and solvent to enhance capacity, exemplified by amine scrubbing where hydrogen sulfide (H₂S) reacts with aqueous amines like monoethanolamine to form non-volatile salts, allowing higher removal efficiencies at lower partial pressures.[43][44] Common equipment for gas-liquid absorption includes packed towers and spray columns, which facilitate intimate contact between counter-current gas and liquid flows to maximize interfacial area. Packed towers use random or structured packing materials to support the liquid film, while spray columns disperse liquid as droplets for simpler, lower-pressure-drop operations suitable for less viscous systems. Design of these units often employs the number of transfer units (NTU) method, where for dilute systems and linear equilibrium with pure inlet solvent (x_{\text{out}} = 0), the overall gas-phase NTU is given by the Colburn formula: \text{NTU}_{OG} = \frac{ \ln \left[ (1 - S) \frac{y_{\text{in}}}{y_{\text{out}}} + S \right] }{1 - S }, where y is the gas mole fraction, S = m \frac{G}{L} is the stripping factor with m as the equilibrium slope (from Henry's law), and G/L as the gas-to-liquid molar flow ratio; the tower height is then Z = \text{HTU} \times \text{NTU}, where HTU is the height of a transfer unit.[45][46] Key applications of gas-liquid absorption include air pollution control, such as the removal of sulfur dioxide (SO₂) from flue gases using alkaline solvents in wet scrubbers to meet emission standards, and natural gas sweetening, where amine-based processes eliminate H₂S and CO₂ to prevent corrosion and meet pipeline specifications. These operations are critical for environmental compliance and resource purification, with packed towers commonly used for high-efficiency SO₂ capture at power plants.[47][48] Energy considerations in gas-liquid absorption primarily involve the heat of absorption, which is the enthalpy change associated with solute dissolution or reaction, affecting process temperature and solvent regeneration. For physical absorption, the heat is typically low and endothermic, while chemical absorption, such as amine-CO₂ reactions, releases significant exothermic heat (around 30-50 kJ/mol for CO₂), necessitating cooling to maintain optimal kinetics and prevent solvent degradation; regeneration often requires heating to reverse the process, contributing to overall energy demands.[49][50]In Biology and Physiology
Nutrient and Water Absorption
In human physiology, nutrient absorption primarily occurs in the small intestine, where structural adaptations such as villi and microvilli dramatically expand the absorptive surface area by factors of 30 to 600 times compared to a smooth tube of equivalent length.[51] These finger-like projections, covered by epithelial cells bearing dense microvilli, facilitate the uptake of carbohydrates, proteins, fats, vitamins, and minerals from digested food.[52] Glucose absorption exemplifies active transport mechanisms, where the sodium-glucose linked transporter 1 (SGLT1) on the apical membrane of enterocytes cotransports one glucose molecule with two sodium ions, powered by the sodium gradient established by the basolateral Na+/K+ ATPase pump.[53] This secondary active transport ensures efficient glucose uptake against its concentration gradient, with subsequent exit via facilitative transporters like GLUT2.[54] Water absorption in the kidneys relies on osmotic gradients and specialized channels to maintain fluid balance, with approximately 180 liters of glomerular filtrate processed daily, of which about 99% is reabsorbed.[55] In the loop of Henle, the countercurrent multiplier system creates a hyperosmotic medullary interstitium, driving passive water reabsorption through aquaporin-1 channels in the descending limb and aquaporin-2 in the collecting ducts under hormonal regulation.[56] This mechanism, combined with active solute transport in the ascending limb, concentrates urine while conserving water essential for homeostasis.[57] In plants, root absorption of water and nutrients supports growth and transpiration, with water entering primarily via passive diffusion through root hairs and aquaporins, then transported upward via xylem vessels driven by root pressure and evaporation.[58] Mineral ions, such as nitrate, require active transport across root cell membranes using proton-coupled symporters, as their soil concentrations are often too low for passive uptake.[59] Mycorrhizal symbioses enhance this process by extending the root system's reach; fungal hyphae increase surface area for mineral acquisition, exchanging them for plant-derived carbohydrates in a mutualistic relationship that improves phosphorus and nitrogen efficiency.[60] Specific nutrients illustrate targeted absorption pathways: dietary iron is absorbed as ferrous ions (Fe²⁺) in the duodenum via the divalent metal transporter 1 (DMT1) on the apical enterocyte membrane, with subsequent export by ferroportin to circulation.[61] Vitamin B12 requires binding to intrinsic factor, a glycoprotein secreted by parietal cells in the stomach, forming a complex that binds to receptors in the ileum for endocytosis and absorption.[62] Disruptions in these processes, as in malabsorption syndromes like celiac disease, lead to deficiencies; autoimmune damage to villi reduces surface area, impairing uptake of iron, folate, and other essentials, often resulting in anemia and osteoporosis.[63] Evolutionary adaptations in nutrient absorption reflect dietary niches, with herbivores developing elongated intestines and microbial fermentation chambers to break down fibrous plant material, enhancing cellulose and hemicellulose utilization.[64] In contrast, carnivores exhibit shorter, more acidic guts optimized for rapid protein and fat digestion, with reduced amylase genes and expanded protease families to efficiently absorb amino acids from animal sources.[65] Post-2010 research highlights the gut microbiome's influence on absorption efficiency, where microbial communities ferment indigestible carbohydrates into short-chain fatty acids, enhancing epithelial integrity and mineral solubility for better uptake.[66] Diverse microbiota also modulate transporter expression, such as increasing SGLT1 activity, thereby improving energy harvest from diet by up to 10% in host-microbe symbioses.[67] Dysbiosis, however, can impair these interactions, reducing nutrient bioavailability as seen in obesity and inflammatory conditions.[68]Drug and Toxin Absorption
In pharmacokinetics, absorption represents the initial phase of the ADME (absorption, distribution, metabolism, and excretion) framework, describing the transfer of a drug from its administration site into the systemic circulation.[5] Bioavailability (F), a key measure of absorption efficiency, is calculated as F = AUC_oral / AUC_IV, where AUC denotes the area under the plasma concentration-time curve for oral and intravenous administration, respectively; this ratio quantifies the fraction of the administered dose that reaches systemic circulation unchanged.[69] Poor absorption can limit therapeutic efficacy, while excessive absorption may increase toxicity risks. Drug absorption occurs via multiple routes, each influenced by physiological barriers. Oral administration, the most common route, involves gastrointestinal uptake but is subject to first-pass metabolism in the liver, where hepatic enzymes metabolize a portion of the drug before systemic distribution, often reducing bioavailability.[70] Transdermal absorption follows Fick's first law of diffusion, expressed as J = -D (dc/dx), where J is the flux, D is the diffusion coefficient, and dc/dx is the concentration gradient across the skin; this passive process allows sustained release but is limited by the stratum corneum barrier.[5] Inhalation provides rapid absorption through the alveolar membrane's large surface area, bypassing first-pass effects and enabling quick onset for respiratory or systemic therapies.[71] Absorption is modulated by physiological and molecular factors. The pH-partition hypothesis posits that non-ionized drug forms predominate in absorption due to their higher lipid solubility, facilitating passive diffusion across membranes; for instance, weak acids absorb better in acidic environments like the stomach.[72] Efflux pumps such as P-glycoprotein (P-gp), an ATP-binding cassette transporter, actively expel drugs from enterocytes back into the intestinal lumen, reducing net absorption and contributing to drug resistance or low bioavailability.[73] For toxins, absorption parallels drug kinetics but often involves unintended exposure. In adults, gastrointestinal absorption of lead is approximately 10-15% of ingested amounts, influenced by dietary factors like calcium deficiency that enhance uptake. Dermal absorption of pesticides varies by compound and formulation; for example, organophosphates like imidacloprid exhibit rates around 5%, posing risks during handling without protective barriers.[74] Pharmacokinetic modeling simplifies absorption prediction using compartmental approaches. In a one-compartment model with first-order absorption and elimination, the plasma concentration is given by the Bateman functionC(t) = \frac{F D k_a}{V (k_a - k_e)} \left( e^{-k_e t} - e^{-k_a t} \right),
where F is bioavailability, D is dose, V is volume of distribution, k_a is the absorption rate constant, k_e is the elimination rate constant, and t is time; this equation assumes first-order kinetics.[75] Clinically, absorption influences bioequivalence testing, where generic drugs must demonstrate comparable AUC and maximum concentration (C_max) to reference products within 80-125% limits to ensure therapeutic similarity.[76] Absorption enhancers like dimethyl sulfoxide (DMSO) improve permeability by disrupting lipid bilayers, aiding delivery of poorly soluble drugs transdermally or orally, though their use requires balancing efficacy against potential irritation.[77]