Nutrient
A nutrient is a chemical substance found in foods that is required by living organisms to sustain basic physiological functions, support growth, repair tissues, and maintain health.[1] In human nutrition, nutrients are broadly classified into six major categories: carbohydrates, lipids (fats), proteins, vitamins, minerals, and water, each playing distinct roles in bodily processes.[1] Certain nutrients, known as essential nutrients, must be obtained through diet because the body cannot synthesize them in sufficient quantities.[2] Nutrients are further divided into macronutrients, needed in larger amounts to provide energy and structural components, and micronutrients, required in smaller quantities to regulate metabolic reactions and prevent deficiencies.[3] Carbohydrates serve as the primary energy source, supplying about 4 kcal per gram and fueling brain function and physical activity.[4] Proteins, composed of amino acids, are vital for building and repairing tissues, enzyme production, and immune function, also providing 4 kcal per gram.[1] Lipids deliver 9 kcal per gram, support cell membrane integrity, hormone synthesis, and nutrient absorption, while water—essential for all cellular activities—comprises about 60% of body weight in adults and facilitates transport, temperature regulation, and chemical reactions.[1] Vitamins and minerals, as micronutrients, do not provide calories but are crucial for enzyme function, antioxidant protection, bone health, and oxygen transport; for instance, vitamin C aids collagen synthesis, while iron is key for hemoglobin production.[5] Deficiencies in any nutrient can lead to specific health issues, such as scurvy from lack of vitamin C or anemia from iron deficiency, underscoring the importance of a balanced diet to meet daily requirements.[5] Overall, nutrients interact synergistically to support homeostasis, with dietary guidelines from authorities like the USDA emphasizing variety to ensure adequate intake across all classes.[6]Overview
Definition
A nutrient is a chemical substance that organisms must obtain from their environment, such as food or other sources, to support essential biological functions including growth, reproduction, maintenance, and energy production. These substances are vital for sustaining life processes across animals, plants, and microorganisms, and they are typically absorbed through digestion or environmental uptake to enable metabolic activities.[7][1] The concept of nutrients emerged in 19th-century chemistry, rooted in advancements in agricultural and organic chemistry. A key milestone was the work of German chemist Justus von Liebig, who in his 1840 publication Die organische Chemie in ihrer Anwendung auf Agrikulturchemie und Physiologie (Organic Chemistry in Its Applications to Agriculture and Physiology) established the foundational principles of mineral nutrition for plants, emphasizing the role of specific chemical elements in growth and challenging prevailing views on soil fertility. Liebig's research shifted understanding from vague notions of "plant food" to precise chemical requirements, laying the groundwork for modern nutritional science.[8][9] Nutrients are categorized based on the quantities required by organisms: macronutrients, needed in larger amounts typically measured in grams per day to provide energy and structural components, and micronutrients, required in trace quantities often in milligrams or micrograms to facilitate regulatory and catalytic processes. This distinction highlights the scale of intake rather than the specific chemical nature, with macronutrients forming the bulk of dietary mass and micronutrients acting in smaller but critical roles.[10][11] Quantification of nutrients varies by type and context, ensuring accurate assessment in nutrition and supplementation. Macronutrients are commonly expressed in grams (g), reflecting their substantial caloric contributions, while micronutrients like minerals are measured in milligrams (mg) or micrograms (mcg), with 1 mg equaling 1,000 mcg. Certain vitamins, particularly fat-soluble ones such as A, D, and E, have historically been quantified using international units (IU), a potency-based measure standardized for biological activity, though regulatory shifts now favor micrograms for precision in labeling.[12][13]Biological Role
Nutrients fulfill three primary biological roles in living organisms: serving as building blocks for the synthesis and repair of tissues and cellular components, providing energy to drive physiological processes, and acting as cofactors or coenzymes that enable enzymatic reactions essential for metabolism. As structural elements, nutrients such as carbohydrates, proteins, and lipids form the foundational macromolecules that constitute cell membranes, organelles, and extracellular matrices, ensuring organismal integrity and function. In their energetic capacity, they are oxidized through cellular respiration or other pathways to generate adenosine triphosphate (ATP), the universal energy currency that powers contraction, transport, and biosynthesis. Furthermore, micronutrients like vitamins and minerals often participate directly in catalysis, stabilizing enzyme active sites or transferring chemical groups during reactions, thereby accelerating biochemical transformations that would otherwise proceed too slowly for life.[14][1][15] These roles are interconnected through metabolic pathways, where nutrients exhibit interdependence to maintain homeostasis and support dynamic cellular activities. Catabolism, the degradative process, breaks down complex nutrient-derived molecules into simpler units, liberating energy and intermediate metabolites that can be redirected for other uses. In contrast, anabolism, the constructive process, harnesses this energy and precursors to assemble macromolecules required for growth and repair, creating a balanced cycle that links breakdown and synthesis. This interplay ensures that nutrient availability modulates the rate and direction of metabolism, preventing imbalances that could impair organismal health.[16][17] The biological imperatives of nutrients extend universally across kingdoms, underpinning survival in animals, plants, fungi, and microbes by facilitating analogous processes of energy acquisition, structural maintenance, and biochemical regulation. In all cases, organisms acquire and process nutrients to adapt to environmental conditions, highlighting a conserved framework for life. Evolutionarily, nutrient requirements have intensified with organismal complexity; primitive single-celled life forms synthesized most necessities de novo, but multicellular eukaryotes developed dependencies on exogenous sources for certain compounds, reflecting adaptations to specialized diets and symbiotic interactions that enhanced efficiency in diverse ecosystems.[18][15][19]Classification of Nutrients
Macronutrients
Macronutrients are nutrients that the body requires in large quantities, typically measured in grams per day, to provide energy, support growth, and maintain structural integrity; they include carbohydrates, proteins, fats, and water.[1] Unlike micronutrients, macronutrients are needed in amounts of 50 grams or more daily for most adults, with dietary guidelines recommending that carbohydrates comprise 45–65% of total caloric intake (approximately 225–325 grams on a 2,000-calorie diet), proteins 10–35% (50–175 grams), and fats 20–35% (44–78 grams).[20] These proportions ensure adequate energy provision, as carbohydrates and proteins each yield about 4 kcal per gram, while fats provide 9 kcal per gram.[1] Water, though not caloric, is classified as a macronutrient due to its intake volume of 2–3 liters per day for adults to support hydration and physiological functions.[21] Carbohydrates are polyhydroxy aldehydes or ketones, or compounds that yield such units upon hydrolysis, primarily composed of carbon, hydrogen, and oxygen in a 1:2:1 ratio.[22] They exist as simple forms like monosaccharides (e.g., glucose and fructose) or complex polysaccharides (e.g., starch and cellulose), serving as the body's preferred energy source by breaking down into glucose for cellular respiration.[22] Each gram of digestible carbohydrate provides approximately 4 kcal of energy, making them essential for brain function and physical activity.[1] In the body, excess carbohydrates are stored as glycogen in liver and muscle cells, while plants store them as starch, highlighting their role in energy reserve.[22] Proteins are large biomolecules composed of one or more long chains of amino acid residues linked by peptide bonds, with 20 standard amino acids forming their primary structure.[23] They perform diverse roles, including acting as enzymes to catalyze biochemical reactions, hormones to regulate physiological processes, and structural components like collagen for tissue support.[24] Dietary proteins are classified as complete if they contain all nine essential amino acids in adequate proportions (typically from animal sources like meat and eggs) or incomplete if lacking one or more (common in plant sources like grains and legumes, which can be complemented by combining foods).[25] Proteins contribute 4 kcal per gram and are vital for tissue repair and immune function.[1] Fats, or lipids, are a heterogeneous group of hydrophobic compounds including triglycerides, phospholipids, and sterols, with fatty acids as their building blocks.[26] They are categorized by saturation: saturated fats lack double bonds between carbon atoms (e.g., in butter), monounsaturated have one double bond (e.g., in olive oil), polyunsaturated have multiple (e.g., in fish oils), and trans fats feature unnatural trans-configured double bonds often from partially hydrogenated oils.[1] Fats are the most energy-dense macronutrient at 9 kcal per gram, providing long-term fuel storage in adipose tissue and aiding in the absorption of fat-soluble vitamins.[1] In the small intestine, dietary fats are emulsified by bile salts and absorbed as micelles, which facilitate the uptake of fatty acids and monoglycerides into enterocytes for repackaging into chylomicrons.[27] Water stands out as a unique macronutrient, constituting about 60% of body weight and essential for nearly all life processes without providing calories.[28] It functions as a solvent for biochemical reactions, a medium for nutrient and waste transport via blood and lymph, and a participant in hydrolysis and other metabolic pathways.[21] Daily requirements vary by age, activity, and climate but generally range from 2 to 3 liters for adults, including water from food and beverages, to maintain fluid balance, regulate temperature, and support digestion.[28] Dehydration impairs these roles, underscoring water's indispensable status alongside other macronutrients.[21]Micronutrients
Micronutrients encompass vitamins and minerals that are essential for human health but required in relatively small quantities, typically ranging from milligrams to micrograms per day, in contrast to the larger amounts needed for macronutrients. These nutrients are broadly classified into two categories: organic compounds known as vitamins and inorganic elements referred to as minerals. Unlike macronutrients, which primarily provide energy, micronutrients support metabolic processes, enzyme function, and cellular maintenance without contributing significantly to caloric intake.[1][29] Vitamins are further divided into fat-soluble and water-soluble types based on their solubility and physiological handling. Fat-soluble vitamins, including A, D, E, and K, are absorbed in the presence of dietary fats and can be stored in the body's adipose tissues and liver, potentially leading to accumulation if intake exceeds needs. Water-soluble vitamins, comprising the B-complex group (such as thiamine, riboflavin, and niacin) and vitamin C, dissolve in water and are not stored extensively; excess amounts are typically excreted in urine, which necessitates regular dietary intake to prevent deficiencies. Stability varies among vitamins, with water-soluble ones often more susceptible to degradation from heat, light, or oxygen exposure during food processing and storage, while fat-soluble vitamins may face bioavailability challenges due to their dependence on lipid digestion. Bioavailability of vitamins is influenced by factors like the food matrix, cooking methods, and individual gut health, with processing sometimes enhancing absorption (e.g., through biofortification) but occasionally reducing it.[30][31][32] Minerals, as inorganic micronutrients, exist primarily in ionic forms within the body and are categorized into major minerals (e.g., calcium, phosphorus, magnesium, sodium, potassium, chloride, and sulfur), which are needed in amounts exceeding 100 mg per day, and trace minerals (e.g., iron, zinc, copper, manganese, iodine, selenium, molybdenum, chromium, and fluoride), required in smaller quantities under 100 mg daily. The body maintains mineral homeostasis largely through renal regulation, where the kidneys filter blood plasma, reabsorb necessary ions in the tubules, and excrete excesses to balance levels and prevent toxicity or deficiency. For instance, calcium and phosphate homeostasis involves parathyroid hormone and vitamin D-mediated adjustments in renal reabsorption.[33][34][35][36] Absorption of micronutrients presents several challenges that can limit their utilization from dietary sources. Gastrointestinal pH plays a key role; for example, an acidic environment in the stomach aids the release and solubility of certain minerals like iron, while alkaline conditions in the intestine may hinder others. Dietary inhibitors such as phytates, found in grains and legumes, bind to minerals like iron, zinc, and calcium, forming insoluble complexes that reduce bioavailability. To address these issues, food fortification strategies—such as adding micronutrients to staples like flour or cereals—have been employed, though they must account for potential interactions with food components to ensure stability and effective delivery.[37][38][39] Micronutrients often exhibit interactions that can be synergistic or antagonistic, affecting their overall absorption and efficacy. Synergies occur when one nutrient enhances the uptake or function of another, such as vitamin C improving non-heme iron absorption by reducing it to a more soluble form. Antagonisms, conversely, involve competition for absorption sites; for example, high calcium intake can inhibit zinc absorption in the intestine due to shared transport mechanisms. These interactions underscore the importance of balanced dietary intake to optimize micronutrient status.[40][39][41]Essentiality
Essential Nutrients
Essential nutrients are chemical substances required for normal physiological function that cannot be synthesized by the body in sufficient quantities to meet its needs, and thus must be obtained through the diet. Their absence or inadequate intake leads to specific deficiency diseases or disorders that impair health and can be reversed by restoring dietary supply. This criterion applies primarily to humans and other animals, distinguishing essential nutrients from non-essential ones that the body can produce endogenously.[1] In humans, essential nutrients are categorized into several classes, including nine essential amino acids, two essential fatty acids (linoleic acid and alpha-linolenic acid), thirteen vitamins, at least eleven minerals, and choline. These span both macronutrients (such as essential amino acids and fatty acids, needed in larger amounts) and micronutrients (such as vitamins and minerals, required in smaller quantities). The exact number of essential minerals can vary slightly based on research, but a common list includes calcium, phosphorus, potassium, sodium, chloride, magnesium, iron, zinc, iodine, copper, and selenium.[1][42] The concept of essential nutrients emerged through historical investigations into deficiency diseases. For instance, in 1747, Scottish physician James Lind conducted the first controlled clinical trial on scurvy, demonstrating that citrus fruits containing vitamin C effectively treated and prevented the condition among sailors, though the nutrient's identity was not confirmed until later. Similarly, in the 1910s, U.S. Public Health Service researcher Joseph Goldberger established pellagra as a dietary deficiency rather than an infectious disease, linking it to poor nutrition; niacin was identified as the key nutrient in 1937. These discoveries underscored the dietary origins of such ailments and paved the way for modern nutrition science.[43][44] Bioavailability—the extent to which essential nutrients are absorbed and utilized—affects their effectiveness and depends on factors like digestion and the gut microbiome. During digestion, enzymes break down food matrices to release nutrients, while the gut microbiota can enhance absorption by producing metabolites or synthesizing certain vitamins (e.g., B vitamins and vitamin K) and influencing mineral uptake through short-chain fatty acid production that lowers intestinal pH. Dysbiosis in the microbiome may reduce bioavailability, exacerbating deficiencies.[45][46] Unlike animals, which require organic forms of essential nutrients such as amino acids, fatty acids, and vitamins due to limited synthetic capabilities, plants primarily need inorganic minerals (e.g., nitrogen, phosphorus, potassium) absorbed from soil, along with carbon dioxide, water, and light for photosynthesis. Plants can synthesize all necessary organic compounds internally, making them autotrophic and obviating the need for dietary intake of complex organics.[15]Conditionally Essential Nutrients
Conditionally essential nutrients, also known as semi-essential nutrients, are substances that the human body can typically synthesize in sufficient quantities under normal conditions but become indispensable when endogenous production is inadequate to meet heightened demands during specific physiological stresses.[47] These conditions arise when metabolic pathways are overwhelmed, such as during rapid growth, injury, or disease, leading to a reliance on dietary intake to prevent deficiencies.[48] The need for conditionally essential nutrients is triggered by various life stages and health challenges, including infancy, pregnancy, aging, and infections, where increased catabolic rates or impaired synthesis exceed the body's capacity. For instance, in pregnancy, demands for amino acids like glycine rise due to fetal development and maternal tissue expansion, potentially rendering them indispensable if protein intake is marginal.[49] In aging populations, reduced synthetic efficiency and chronic inflammation can elevate requirements for nutrients like carnitine, while infections such as sepsis accelerate glutamine depletion through immune activation and tissue repair processes.[50] These mechanisms highlight how conditional essentiality stems from an imbalance between synthesis and utilization rather than absolute inability to produce the nutrient. Prominent examples include arginine, which is conditionally essential in neonates due to immature urea cycle function and in adults during wound healing, where it supports nitric oxide production for vasodilation and collagen synthesis.[51] Glutamine similarly becomes critical in sepsis and trauma, as hypercatabolism depletes muscle stores, impairing immune function and gut barrier integrity if not supplemented.[52] Carnitine exemplifies this in genetic disorders like primary carnitine deficiency, where transport defects necessitate exogenous supply to facilitate fatty acid oxidation and prevent metabolic crises.[53] Clinically, supplementation of conditionally essential nutrients is guided by evidence-based protocols, particularly in critical care settings; for example, intravenous glutamine is recommended for select patients with trauma or burns to mitigate infection risk and support recovery, though high doses require caution to avoid adverse effects.[54] Post-2020 research has underscored their roles in COVID-19 recovery, with studies showing glutamine deficiency correlating with disease severity and arginine supplementation potentially aiding immune modulation and endothelial function in severe cases.[55] In long COVID contexts, addressing conditional deficits in amino acids like these through targeted nutrition has been linked to improved inflammatory profiles and reduced sequelae.[56]Specific Essential Nutrients
Amino Acids
Amino acids are organic compounds that serve as the building blocks of proteins, and among the 20 standard amino acids used in human protein synthesis, nine are classified as essential because the human body cannot synthesize them and must obtain them from the diet. These essential amino acids are histidine (C₆H₉N₃O₂), isoleucine (C₆H₁₃NO₂), leucine (C₆H₁₃NO₂), lysine (C₆H₁₄N₂O₂), methionine (C₅H₁₁NO₂S), phenylalanine (C₉H₁₁NO₂), threonine (C₄H₉NO₃), tryptophan (C₁₁H₁₂N₂O₂), and valine (C₅H₁₁NO₂).[23][57] Essential amino acids play critical roles in protein synthesis, where they are incorporated into polypeptide chains to form functional proteins essential for growth, repair, and maintenance of tissues. Beyond structural roles, certain essential amino acids act as precursors for bioactive molecules; for instance, tryptophan serves as the sole precursor for the neurotransmitter serotonin, which regulates mood, appetite, and sleep.[23][58] Branched-chain amino acids (BCAAs)—isoleucine, leucine, and valine—are particularly important for muscle repair and metabolism, as they stimulate muscle protein synthesis and provide energy during prolonged exercise by being oxidized in skeletal muscle.[59][60] Dietary sources of essential amino acids are primarily proteins, with animal-based foods such as meat, eggs, dairy, and fish providing complete proteins that contain all nine in adequate proportions. Plant-based sources often lack one or more essential amino acids in sufficient amounts—for example, grains like rice are low in lysine—but combining complementary plant proteins, such as rice with beans, can yield a complete amino acid profile by balancing deficiencies.[61][62] Deficiency in essential amino acids, typically arising from inadequate protein intake, can lead to impaired protein synthesis and symptoms resembling kwashiorkor, including edema, fatty liver, and skin lesions due to hypoalbuminemia and metabolic disruptions.[63][64] Recent research has highlighted leucine's specific role in activating the mechanistic target of rapamycin complex 1 (mTORC1) signaling pathway, which regulates protein synthesis and cellular growth; in the context of aging, leucine supplementation has been shown to mitigate age-related muscle loss (sarcopenia) by enhancing mTORC1-mediated anabolism and autophagy balance.[65][66]Fatty Acids
Fatty acids become essential nutrients when the human body cannot synthesize them in sufficient quantities, necessitating dietary intake to prevent deficiency. The two primary essential fatty acids are linoleic acid (LA), an omega-6 polyunsaturated fatty acid denoted as 18:2 n-6 (an 18-carbon chain with double bonds starting at the sixth carbon from the methyl end), and alpha-linolenic acid (ALA), an omega-3 polyunsaturated fatty acid denoted as 18:3 n-3 (an 18-carbon chain with double bonds starting at the third carbon from the methyl end).[67][68] These polyunsaturated structures, characterized by multiple cis double bonds, distinguish them from saturated or monounsaturated fatty acids and enable their incorporation into cell membranes.[67] Linoleic acid and alpha-linolenic acid serve critical biochemical roles, primarily as precursors for longer-chain polyunsaturated fatty acids involved in cellular signaling and structural integrity. LA is elongated and desaturated to form arachidonic acid (AA, 20:4 n-6), which acts as a substrate for eicosanoids such as prostaglandins, thromboxanes, and leukotrienes that regulate inflammation, blood clotting, and immune responses.[67] Similarly, ALA is converted to eicosapentaenoic acid (EPA, 20:5 n-3) and docosahexaenoic acid (DHA, 22:6 n-3), which compete with AA-derived eicosanoids to produce less inflammatory mediators, thereby exerting anti-inflammatory effects.[67] Both LA and ALA contribute to membrane fluidity by integrating into phospholipid bilayers, enhancing membrane dynamics essential for receptor function and signal transduction.[67] Dietary sources of linoleic acid include vegetable oils such as soybean, corn, sunflower, and safflower oils, which are rich in this omega-6 fatty acid.[67] Alpha-linolenic acid is primarily found in plant-based sources like flaxseed oil, canola oil, soybean oil, chia seeds, and walnuts.[67] The body's conversion of ALA to the more bioactive EPA and DHA is inefficient, with estimates indicating less than 5% conversion to EPA and under 0.5% to DHA in adults, particularly lower in men due to hormonal differences.[67] Arachidonic acid and DHA are considered conditionally essential in specific contexts, such as infancy, pregnancy, or conditions impairing endogenous synthesis, where direct dietary provision from animal sources like meat, eggs, and fatty fish supports optimal brain and retinal development.[69] Recent guidelines underscore the cardiovascular benefits of omega-3 fatty acids, with the American Heart Association's 2023 update recommending consumption of fatty fish at least twice weekly to achieve adequate EPA and DHA intake for reducing risks of heart disease, stroke, and inflammation-related conditions.[70][71] This aligns with evidence that higher omega-3 levels from diet or supplements can lower triglycerides and improve endothelial function, emphasizing the need for balanced intake to maintain essential fatty acid homeostasis.[70]Vitamins
Vitamins are organic compounds required in small amounts for normal growth, reproduction, and maintenance of health, classified as a subset of micronutrients that cannot be synthesized adequately by the human body and must be obtained from the diet.[1] They are divided into two main groups based on solubility: fat-soluble vitamins (A, D, E, and K), which are absorbed along with dietary fats and stored in the body's fatty tissues and liver, and water-soluble vitamins (B-complex and C), which are absorbed directly into the bloodstream and excreted in urine if in excess.[72] Fat-soluble vitamins require bile salts and pancreatic enzymes for incorporation into micelles in the small intestine, enhancing their absorption when consumed with fats, whereas water-soluble vitamins are more prone to degradation from heat, light, and processing, leading to potential losses during food preparation.[31][30] The fat-soluble vitamins include vitamin A (retinol and its esters, with provitamin forms like beta-carotene), which functions in vision maintenance, immune response, and epithelial cell integrity by supporting rhodopsin formation in the retina.[73] Vitamin D (calciferol, including ergocalciferol [D2] and cholecalciferol [D3]) regulates calcium and phosphorus homeostasis for bone mineralization and has emerged in post-2020 studies as a modulator of immune function, reducing inflammation and enhancing antimicrobial peptides during respiratory infections like COVID-19.[74][75] Vitamin E (tocopherols and tocotrienols, with alpha-tocopherol being most active) acts as an antioxidant, protecting cell membranes from oxidative damage by neutralizing free radicals.[31] Vitamin K (phylloquinone [K1] from plants and menaquinones [K2] from bacteria) is essential for blood clotting through gamma-carboxylation of proteins like prothrombin and for bone health via osteocalcin activation.[76] These vitamins are measured in international units (IU) or milligrams (mg), with recommended dietary allowances (RDAs) varying by age and life stage; for example, adult RDA for vitamin A is 900 mcg retinol activity equivalents (RAE), where 1 RAE equals 12 mcg beta-carotene due to its lower conversion efficiency.[73] The water-soluble vitamins encompass the B-complex group and vitamin C. Vitamin B1 (thiamine, active as thiamine pyrophosphate [TPP]) serves as a cofactor in carbohydrate metabolism and nerve function, with historical deficiency causing beriberi, a condition of neuropathy and heart failure prevalent in polished rice-dependent diets in the early 20th century.[30][77] Vitamin B2 (riboflavin, forming flavin adenine dinucleotide [FAD] and flavin mononucleotide [FMN]) participates in redox reactions for energy production and antioxidant defense. Vitamin B3 (niacin, as nicotinamide adenine dinucleotide [NAD]) is crucial for electron transport in metabolism and DNA repair. Vitamin B5 (pantothenic acid, component of coenzyme A [CoA]) aids in fatty acid synthesis and energy derivation from food. Vitamin B6 (pyridoxine and derivatives like pyridoxal phosphate) facilitates amino acid metabolism, neurotransmitter synthesis, and hemoglobin production. Vitamin B7 (biotin) acts as a cofactor for carboxylase enzymes in gluconeogenesis and fatty acid synthesis. Vitamin B9 (folate, as tetrahydrofolate [THF]) supports DNA synthesis and methylation reactions critical for cell division. Vitamin B12 (cobalamin, including methylcobalamin and adenosylcobalamin) is involved in myelin sheath maintenance, red blood cell formation, and one-carbon metabolism. Finally, vitamin C (ascorbic acid) functions in collagen synthesis, iron absorption, and as an antioxidant, with various vitamers differing in bioavailability; RDAs are expressed in mg, such as 90 mg/day for adult men for vitamin C.[30][72] Vitamin D deficiency has historically led to rickets, characterized by softened bones in children due to impaired mineralization, as observed in industrialized populations with limited sunlight exposure before supplementation programs.[77]Minerals
Minerals are inorganic elements essential to human physiology, serving as cofactors in enzymatic reactions, components of structural tissues, and regulators of fluid balance and nerve function. Unlike organic nutrients such as vitamins, minerals exist primarily as ions and are not synthesized by the body, necessitating dietary intake. They are categorized into macrominerals, required in quantities exceeding 100 mg daily for adults, and trace minerals, needed in amounts under 100 mg daily but critical for specialized processes. There are over 15 recognized essential minerals, with roles in homeostasis tightly regulated to prevent deficiencies or excesses that can lead to disorders like osteoporosis or anemia.[78][79]Macrominerals
Macrominerals form the bulk of the body's mineral content and support foundational physiological functions. Calcium, the most abundant mineral in the human body, is vital for bone and tooth mineralization, muscle contraction, nerve signaling, and blood clotting, with approximately 99% stored in skeletal tissues; the recommended dietary allowance (RDA) for adults is 1,000–1,200 mg per day.[80][81] Phosphorus works synergistically with calcium in bone formation and is a key constituent of ATP, nucleic acids, and phospholipids, essential for energy transfer; its RDA is 700 mg daily for adults.[82][83] Magnesium activates over 300 enzymes involved in ATP synthesis, protein synthesis, and neuromuscular function, with RDAs ranging from 310 mg for women to 420 mg for men aged 31 and older.[84][83] Sodium and potassium act as major electrolytes, maintaining cellular fluid balance, membrane potential, and acid-base equilibrium; sodium's adequate intake is 1,500 mg daily, while potassium's is 4,700 mg, with imbalances linked to hypertension.[1][85] Chloride, often paired with sodium in salts, supports gastric acid production and osmotic balance, while sulfur, derived from dietary proteins, contributes to the structure of amino acids like cysteine and methionine, aiding detoxification and antioxidant activity.[86][87]Trace Minerals
Trace minerals, though required in minute quantities, are indispensable for enzymatic catalysis, antioxidant defense, and hormone synthesis. Iron is central to hemoglobin and myoglobin for oxygen transport and storage, as well as cytochrome enzymes in energy metabolism; adult RDAs are 8 mg for men and 18 mg for premenopausal women.[88][1] Zinc facilitates DNA and RNA synthesis, cell division, immune response, and wound healing through its role in over 300 enzymes; RDAs are 11 mg for men and 8 mg for women.[89] Copper functions as a cofactor in superoxide dismutase, an antioxidant enzyme that protects cells from oxidative damage, and in iron metabolism; the RDA is 900 mcg daily. Manganese supports bone development, metabolism of carbohydrates and cholesterol, and antioxidant defenses via enzymes like superoxide dismutase; adequate intake is 2.3 mg for men and 1.8 mg for women.[87] Iodine is incorporated into thyroid hormones thyroxine and triiodothyronine, regulating metabolism, growth, and development; the RDA is 150 mcg for adults.[90] Selenium is a component of selenoproteins, including glutathione peroxidase, which neutralizes reactive oxygen species; the RDA is 55 mcg.[87] Molybdenum acts as a cofactor for enzymes involved in sulfur amino acid metabolism and detoxification of sulfites; the RDA is 45 mcg.[87] Chromium enhances insulin action and glucose metabolism; adequate intake is 35 mcg for men and 25 mcg for women.[87] Fluoride, with debated essentiality, strengthens tooth enamel and may support bone health; adequate intake is 4 mg for men and 3 mg for women, primarily from fluoridated water.[86][87] Mineral absorption occurs mainly in the small intestine via active and passive transport mechanisms, often enhanced by chelation with organic ligands like amino acids or peptides to improve solubility and prevent precipitation in the alkaline environment. Specific transport proteins facilitate uptake: for instance, divalent metal transporter 1 (DMT1) handles iron and other divalent cations, while zinc is absorbed via ZIP transporters.[91][92] Once absorbed, minerals bind to carrier proteins for distribution, such as transferrin for iron or albumin for calcium, ensuring targeted delivery to tissues.[93] Homeostasis of minerals involves hormonal and renal regulation to maintain plasma concentrations within narrow limits. For calcium, the parathyroid glands secrete parathyroid hormone (PTH) in response to low serum levels, stimulating renal reabsorption in the distal tubules, bone resorption, and intestinal absorption via vitamin D activation.[94][95] Sodium homeostasis is primarily managed by the kidneys, where filtration and reabsorption—regulated by aldosterone, atrial natriuretic peptide, and renal sympathetic nerves—adjust excretion to match intake, preventing volume depletion or overload.[96][97] Dietary sources vary by mineral, influencing bioavailability. Dairy products like milk and yogurt are primary sources of calcium, providing 300 mg per cup of milk. Seafood, including fish and shellfish, supplies iodine (up to 100 mcg per serving of cod) and selenium (e.g., 40 mcg in tuna). Certain compounds, such as goitrogens in cruciferous vegetables like cabbage and broccoli, can inhibit iodine absorption by competing with iodide for uptake at the sodium-iodide symporter in the thyroid.[98][90] In a 2024 update, the World Health Organization revised haemoglobin concentration cutoffs to define anaemia, facilitating improved detection and intervention strategies, including iron fortification of staple foods, to address global iron deficiency anaemia affecting over 1.9 billion people.[99][100]Choline
Choline is a water-soluble quaternary ammonium compound with the chemical formula C₅H₁₄NO⁺, characterized by a positively charged nitrogen atom bonded to three methyl groups and a hydroxylethyl group.[101] Although the human body can synthesize small amounts endogenously, primarily in the liver, choline is considered a quasi-essential nutrient because dietary intake is necessary for most individuals to meet physiological demands, particularly those with genetic variations such as polymorphisms in the phosphatidylethanolamine N-methyltransferase (PEMT) gene that impair synthesis.[102] These variants increase reliance on dietary sources, making choline essential for affected populations to prevent metabolic disruptions.[102] Choline serves critical roles in cellular processes, acting as a precursor for several key biomolecules involved in cell signaling and structure. It is acetylated to form acetylcholine, a neurotransmitter essential for nerve impulse transmission, muscle contraction, memory, and mood regulation.[42] Additionally, choline is incorporated into phosphatidylcholine, a major phospholipid component of cell membranes that maintains structural integrity and facilitates lipid transport.[102] Through oxidation, choline is converted to betaine, which functions as a methyl group donor in one-carbon metabolism, supporting DNA methylation and homocysteine remethylation.[103] Dietary sources of choline are abundant in animal and plant foods, with eggs, beef liver, and soybeans among the richest providers; for instance, a single large egg yolk contains approximately 147 mg, while 3 ounces of beef liver provides about 356 mg.[42] Endogenous synthesis occurs via the PEMT enzyme in the liver, which methylates phosphatidylethanolamine to produce phosphatidylcholine, but this pathway is limited during pregnancy due to heightened demands for fetal development and maternal adaptations, often rendering dietary intake insufficient.[104] The Adequate Intake (AI) for choline, established by the National Academies of Sciences, Engineering, and Medicine, is 550 mg per day for adult men and 425 mg per day for adult women, with higher levels recommended during pregnancy (450 mg/day) and lactation (550 mg/day) to support increased needs.[42] Deficiency, which can manifest as fatty liver disease due to impaired lipid export from hepatocytes, is particularly prevalent in conditions like chronic liver disease where PEMT activity is compromised.[42] Recent research, including a 2022 prospective cohort study of over 3,000 older adults, has linked low dietary choline intake to elevated risks of dementia and Alzheimer's disease, highlighting its neuroprotective potential through sustained acetylcholine production and membrane maintenance.[105]Functions and Metabolism
Energy Production
Nutrients serve as primary substrates for ATP generation through catabolic pathways that break down carbohydrates, fats, and proteins into high-energy molecules. Carbohydrates, primarily in the form of glucose, undergo glycolysis in the cytoplasm, converting one molecule of glucose to two molecules of pyruvate while yielding a net of 2 ATP and 2 NADH.[106] This process is anaerobic and rapid, providing quick energy for cellular needs. Fats are metabolized via beta-oxidation in the mitochondria, where fatty acyl-CoA chains are sequentially shortened by two carbons, producing acetyl-CoA, NADH, and FADH2 for entry into the electron transport chain.[107] Proteins contribute through amino acid catabolism, where the carbon skeletons of glucogenic and ketogenic amino acids are funneled into gluconeogenesis, the Krebs cycle, or ketone body formation, with nitrogen waste handled by the urea cycle to enable energy extraction without toxicity.[108] The complete aerobic oxidation of glucose integrates glycolysis with the Krebs cycle and oxidative phosphorylation, summarized by the equation: \text{C}_6\text{H}_{12}\text{O}_6 + 6\text{O}_2 \rightarrow 6\text{CO}_2 + 6\text{H}_2\text{O} + \sim 30-38 \text{ ATP} This yields approximately 30-38 ATP per glucose molecule, depending on shuttle efficiencies and proton leak.[109] For fats, beta-oxidation of palmitate (a 16-carbon fatty acid) generates 8 acetyl-CoA units, along with 7 NADH and 7 FADH2, resulting in a net yield of approximately 106 ATP after accounting for activation costs.[110] These pathways converge on the Krebs cycle, where acetyl-CoA is oxidized to CO2, producing additional reducing equivalents for ATP synthesis. Several vitamins and minerals act as essential cofactors in these processes. B-vitamins, such as thiamine (vitamin B1) as thiamine pyrophosphate, are critical for the pyruvate dehydrogenase complex, linking glycolysis to the Krebs cycle by decarboxylating pyruvate to acetyl-CoA.[111] Other B-vitamins like riboflavin (FAD) and niacin (NAD) facilitate electron transfer in beta-oxidation and the Krebs cycle. Magnesium, a key mineral, stabilizes ATP in ATPase enzymes and supports over 300 reactions, including those in glycolysis and oxidative phosphorylation for efficient energy transfer.[112] The efficiency of energy production varies by nutrient: carbohydrate oxidation achieves about 40% thermodynamic efficiency in converting chemical energy to ATP, while fats provide higher caloric density (9 kcal/g versus 4 kcal/g for carbohydrates) but slower mobilization due to the multi-step beta-oxidation process.[113][114] In conditions like diabetes, dysregulation impairs glucose utilization, reducing glycolysis and aerobic ATP yield from carbohydrates, shifting reliance to fats and leading to metabolic inefficiency and complications.[115]Structural and Regulatory Roles
Nutrients play essential structural roles in biological systems by forming the foundational components of cellular and tissue architecture. Amino acids serve as the primary building blocks of proteins, which provide structural support to cells, tissues, and organs, including the formation of collagen in connective tissues and keratin in hair and nails.[23] These proteins also contribute to the structural organization of DNA through histone complexes, where specific amino acids facilitate chromatin packaging and stability. Fatty acids, particularly those incorporated into phospholipids, form the lipid bilayer of cell membranes, enabling compartmentalization, fluidity, and selective permeability essential for cellular integrity. Minerals such as calcium and phosphate are critical for skeletal structure, combining to form hydroxyapatite crystals with the formula \ce{Ca10(PO4)6(OH)2}, which constitute approximately 70% of bone mineral content and provide rigidity and strength to the skeletal system.[116][117] Beyond structure, nutrients exert regulatory functions by modulating biochemical pathways, gene expression, and physiological homeostasis. Vitamins often act as coenzymes in enzymatic reactions; for instance, vitamin B6 in its active form, pyridoxal phosphate, facilitates transamination reactions that transfer amino groups between molecules, aiding in amino acid metabolism and neurotransmitter synthesis. Iodine is indispensable for the biosynthesis of thyroid hormones, such as thyroxine, where it is incorporated into the tyrosine residues of thyroglobulin to regulate metabolism, growth, and development. Antioxidants like vitamin E (α-tocopherol) protect cell membranes from oxidative damage by scavenging lipid peroxyl radicals, while selenium, as a component of glutathione peroxidase, neutralizes reactive oxygen species to maintain redox balance and prevent cellular injury.[118][119][120] Nutrients also influence gene expression and long-term cellular regulation. Folate provides one-carbon units for DNA methylation, an epigenetic modification that silences genes by adding methyl groups to cytosine residues, thereby controlling developmental processes and disease susceptibility. Zinc enables the formation of zinc finger motifs in transcription factors, which bind specific DNA sequences to activate or repress gene transcription, playing key roles in immune response and cell differentiation. In metabolic homeostasis, amino acids derived from dietary proteins stimulate the synthesis of peptide hormones like insulin and glucagon in pancreatic β- and α-cells, respectively; these hormones form a feedback loop to regulate blood glucose levels and amino acid catabolism, preventing metabolic imbalances.[121][122][123] Links between nutrients and aging involve mechanisms of genomic stability. Folate and vitamin B12 support telomere maintenance by ensuring proper DNA synthesis and repair; lower levels of these vitamins are associated with shortened telomeres in women, potentially accelerating cellular senescence and contributing to age-related decline.[124]Sources and Requirements
Dietary and Environmental Sources
Nutrients are primarily obtained through dietary sources, which can be categorized into animal- and plant-based foods, each providing distinct profiles of essential macronutrients and micronutrients with varying bioavailability. Animal-sourced foods, such as meat, poultry, eggs, fish, and dairy products, are rich in high-quality protein, essential amino acids, vitamin B12, calcium, and vitamin D. For instance, red meat and organ meats serve as primary sources of vitamin B12, which is almost exclusively found in animal products due to its synthesis by bacteria in animal rumens or guts. Dairy products like milk and cheese contribute significant amounts of bioavailable calcium and vitamin D, supporting bone health when consumed regularly.[125] Plant-based foods complement animal sources by supplying carbohydrates, B vitamins, fiber, iron, and vitamin C, though bioavailability often depends on preparation methods and co-consumption with enhancers. Grains such as wheat, rice, and oats provide complex carbohydrates and B vitamins like thiamin and niacin, forming the staple energy base in many diets. Legumes, including beans, lentils, and peas, offer dietary fiber and non-heme iron, with the latter being more absorbable when paired with vitamin C-rich foods. Fruits like citrus, berries, and kiwi are key sources of vitamin C, an antioxidant that also aids iron absorption from plant sources.[126][127] Beyond diet, environmental sources play a critical role in nutrient acquisition, particularly through natural processes. Sunlight exposure triggers the synthesis of vitamin D in the skin via the conversion of 7-dehydrocholesterol to previtamin D3 upon UVB radiation absorption, accounting for up to 90% of vitamin D needs in sun-exposed individuals. Soil microbes contribute to nutrient availability in plants, facilitating the cycling of nitrogen, phosphorus, and other elements that ultimately enter the human food chain through agriculture, though human direct uptake is limited to fortified or supplemented forms.[128][129][130] Food fortification and enrichment enhance nutrient intake by adding essential micronutrients to commonly consumed staples, addressing gaps in natural sources. Universal salt iodization introduces iodine to prevent deficiencies, while rice fortification with vitamin A—achieved by coating grains with micronutrient premixes—targets populations reliant on rice as a dietary staple. These interventions improve bioavailability without altering food's sensory qualities.[131][132] Bioavailability of certain nutrients can be optimized through dietary synergies, such as consuming vitamin C alongside non-heme iron from plants, where ascorbic acid reduces ferric iron to its more absorbable ferrous form and chelates it for better uptake. This interaction is dose-dependent and can counteract inhibitors like phytates in grains.[133][134] Global dietary variations influence nutrient sourcing, with developing regions often featuring diets heavy in rice or maize that may lack sufficient protein, iron, vitamin A, and iodine due to limited access to diverse animal and fortified foods. In low- and middle-income countries, such monotonous staples contribute to widespread micronutrient deficiencies affecting over 2 billion people, particularly children and pregnant women.[135][136]Recommended Intakes and Guidelines
Recommended Dietary Allowances (RDAs) and Adequate Intakes (AIs) represent the levels of nutrient intake sufficient to meet the requirements of nearly all (97-98%) healthy individuals in a specific life stage and gender group, while Tolerable Upper Intake Levels (ULs) indicate the maximum daily intake unlikely to cause adverse effects. These values are established by authoritative bodies such as the National Academies of Sciences, Engineering, and Medicine (formerly the Institute of Medicine, IOM) in the United States through Dietary Reference Intakes (DRIs), the European Food Safety Authority (EFSA) via Dietary Reference Values (DRVs), and the World Health Organization (WHO) for global standards.[137][138][139] Recommendations vary by factors including age, sex, physical activity level, and physiological state, with higher needs often required during growth, reproduction, or increased metabolic demands. For instance, the IOM sets the RDA for protein at 0.8 g per kg of body weight for adults, equating to about 56 g/day for a 70 kg man or 46 g/day for a 57 kg woman, though active individuals may require up to 1.2-2.0 g/kg based on evidence from exercise physiology studies.[140] Vitamin C RDA is 90 mg/day for adult men and 75 mg/day for women, with smokers needing an additional 35 mg/day to account for oxidative stress.[127] EFSA's Population Reference Intake (PRI) for vitamin C aligns closely at 110 mg/day for men and 95 mg/day for women, derived from factorial methods estimating antioxidant needs.[141] WHO guidelines emphasize intakes to prevent deficiencies, such as at least 45 mg/day for vitamin C in adults to support immune function in low-resource settings.[142] Life stage adjustments reflect elevated demands; for example, the IOM RDA for folate increases to 600 mcg dietary folate equivalents (DFE) during pregnancy to support fetal neural tube development, compared to 400 mcg DFE for non-pregnant women of childbearing age.[143] Iron requirements are particularly high in infancy and pregnancy, with the IOM RDA at 11 mg/day for infants aged 7-12 months to replenish stores post-birth, and 27 mg/day for pregnant women to prevent maternal anemia.[88] EFSA sets the PRI for iron at 11 mg/day for adult men and postmenopausal women, but 16 mg/day for premenopausal women due to menstrual losses, with average requirements during pregnancy estimated at 9.9 mg/day adjusted for bioavailability.[138] ULs prevent toxicity, such as 45 mg/day for iron in adults to avoid gastrointestinal distress, and 2,000 mg/day for vitamin C to limit osmotic diarrhea.[88][127] Nutrient status assessment combines dietary methods like 24-hour recalls or food frequency questionnaires with biomarkers for precision. Dietary recalls capture usual intake but are subject to reporting bias, while biomarkers such as serum 25-hydroxyvitamin D (25-OH D) provide objective measures of vitamin D status, with levels below 30 nmol/L indicating deficiency per IOM criteria.[144] For iron, serum ferritin below 15 mcg/L signals depleted stores, complementing intake data.[88] The Scientific Report of the 2025 Dietary Guidelines Advisory Committee (published December 2024) emphasizes seafood consumption (8-10.5 oz equivalents per week) as a source of eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA) to support cardiovascular health in dietary patterns, aligning with prior recommendations. The IOM Adequate Intakes for alpha-linolenic acid (ALA), the primary plant-based omega-3, remain at 1.6 g/day for men and 1.1 g/day for women, with organizations suggesting 250-500 mg/day combined EPA and DHA from seafood for heart health benefits. The final Dietary Guidelines for Americans 2025-2030 are expected by the end of 2025.[145][67]| Nutrient | IOM RDA/AI (Adults 19-50 years) | Life Stage Example (IOM) | UL (Adults, IOM) | EFSA PRI Example (Adults) |
|---|---|---|---|---|
| Protein | 0.8 g/kg body weight | Pregnancy: +25 g/day | None established | 0.83 g/kg body weight |
| Vitamin C | Men: 90 mg/day; Women: 75 mg/day | Pregnancy: +10 mg/day | 2,000 mg/day | Men: 110 mg/day; Women: 95 mg/day |
| Folate | 400 mcg DFE/day | Pregnancy: 600 mcg DFE/day | 1,000 mcg/day (synthetic) | 330 mcg/day |
| Iron | Men: 8 mg/day; Women: 18 mg/day | Infants 7-12 mo: 11 mg/day | 45 mg/day | Men: 11 mg/day; Women: 16 mg/day |
| Omega-3 (ALA) | Men: 1.6 g/day AI; Women: 1.1 g/day AI | Pregnancy: +0.2 g/day | 3 g/day (EPA/DHA) | 2% of energy intake |