Partial melting
Partial melting is a geological process in which rocks or mineral assemblages melt incompletely over a temperature range, producing a liquid melt (magma) with a composition distinct from the original solid material, typically enriched in elements that are incompatible with early-forming crystals.[1][2] This occurs because individual minerals within the rock have different melting points, with lower-temperature phases melting first, resulting in a melt that is often more silica-rich than the bulk source.[3] Unlike complete melting, partial melting generates heterogeneous mixtures of melt, unmelted solids, and volatiles, which can segregate to form magmas that ascend through the crust.[1] The primary mechanisms driving partial melting include decompression, flux addition, and conductive heating. Decompression melting happens when hot mantle rock rises toward lower pressures, such as at mid-ocean ridges or hotspots, causing the melting point to drop and intersect the geothermal gradient.[1] Flux melting, or fluid-induced melting, occurs when water or carbon dioxide is introduced, lowering the solidus temperature; this is common in subduction zones where hydrous fluids from the subducting slab infiltrate the overlying mantle.[3] Conductive heating, though less common, involves heat transfer from an intruding magma body raising temperatures in surrounding rocks to initiate melting, often in continental crust.[1] Partial melting is essential for igneous petrogenesis, as the degree of melting (typically 1–20% in natural settings) and source rock composition determine magma types, from mafic basalts derived from mantle peridotite to felsic rhyolites from crustal sources.[1] For instance, higher-degree partial melting (typically 10–25%) of the mantle at divergent boundaries produces primitive basaltic magmas.[3][4] This process influences volcanic activity, crustal evolution, and the geochemical diversity of Earth's igneous rocks, with implications for plate tectonics and planetary differentiation.[1]Fundamentals
Definition and Process
Partial melting is the geological process in which only a portion of a rock transforms into liquid (melt or magma), leaving behind a solid crystalline residue, typically occurring in the Earth's mantle or crust when temperatures exceed the solidus but remain below the liquidus under specific pressure conditions. This process is essential for generating diverse igneous magmas and facilitating crustal differentiation, as the melt often segregates from the residue and ascends to form volcanic or plutonic rocks.[5] The fundamental mechanism of partial melting is incongruent melting, where individual minerals melt at different temperatures due to their varying melting points, causing the initial melt to be enriched in components from the lower-melting-point phases—such as silica or alkalis—resulting in a composition distinct from the bulk parent rock. For instance, in a typical mantle peridotite, olivine and pyroxenes may begin melting before more refractory minerals like garnet, producing a basaltic melt that is more silica-rich relative to the solid residue. In contrast, complete (congruent) melting, where the entire rock liquefies uniformly to match the bulk composition, is uncommon in natural geological settings because rocks are polymineralic aggregates with heterogeneous phase equilibria.[5] The extent of partial melting, expressed as the melt fraction (typically 1–20% by volume), critically controls the volume of magma generated and its geochemical signature; low degrees (e.g., 1–5%) yield small volumes of highly incompatible-element-enriched melts, while higher degrees (10–20%) produce larger volumes closer to the bulk composition. This variability arises from the interplay of temperature exceeding the solidus and other factors like pressure, though the core process remains driven by selective mineral melting.[5] The concept of partial melting emerged in the early 20th century through petrological analyses of igneous rock suites, with Norman L. Bowen providing foundational insights in his 1928 treatise on magmatic evolution, where he described fractional melting as the reverse of fractional crystallization to explain compositional diversity in igneous rocks. Significant progress followed in the 1960s via experimental petrology, as high-pressure apparatus enabled direct observation of partial melting in mantle-like compositions, exemplified by Ikuo Kushiro's experiments demonstrating melt generation from hydrous peridotites under upper mantle conditions.[6]79[1685:EOWOTM]2.0.CO;2)Equilibrium vs. Fractional Melting
In equilibrium melting, also known as batch melting, the melt generated during partial melting remains in continuous contact with the solid residue, allowing for ongoing chemical reactions that maintain equilibrium between the liquid and solid phases throughout the process.[2] This closed-system behavior results in the melt composition evolving gradually with the degree of melting, remaining chemically tied to the residue as incompatible elements partition into the liquid while compatible elements concentrate in the solids.[7] Consequently, the overall composition of the system stays representative of the original source rock, producing a spectrum of melt compositions that reflect varying extents of melting.[8] In contrast, fractional melting involves the immediate extraction and segregation of the melt from the residue as soon as it forms, preventing further reaction between the liquid and solids.[2] This dynamic process leads to progressive depletion of the source rock, particularly of incompatible elements, with early-formed melts being more evolved and enriched in silica or other fractionated components compared to later increments.[8] Schematically, equilibrium melting can be visualized as a single batch where the melt and residue coexist like ingredients stirred in a pot, yielding a blended product, whereas fractional melting resembles incrementally siphoning off portions of liquid from the mixture, leaving behind a progressively altered solid.[7] The key distinction lies in their applicability to geological settings: equilibrium melting is best suited to closed-system models with limited melt volumes, such as localized diapiric upwelling where segregation is inefficient, while fractional melting predominates in open, dynamic systems like mid-ocean ridge mantle upwelling, where buoyancy-driven porous flow facilitates rapid melt extraction.[7] Experimental phase equilibria studies, such as those on peridotite systems, demonstrate that fractional melting produces lower overall melt fractions under identical conditions compared to equilibrium melting, as each extracted increment raises the solidus temperature of the residue, requiring additional heat for further melting.[2] This evidence underscores fractional melting's role in generating diverse magma compositions observed in natural settings, though equilibrium processes may dominate when source rock composition inhibits efficient segregation.[8]Influencing Parameters
Source Rock Composition
The composition of the source rock fundamentally governs the initiation, degree, and chemical characteristics of partial melting in geological systems. In multi-component silicate rocks, melting does not occur uniformly but follows eutectic-like behavior, where the first melts form at the lowest temperature corresponding to the eutectic composition in the phase diagram, often differing from the bulk source modal mineralogy. This process is exemplified in non-modal melting, where minerals dissolve into the melt in proportions that deviate from their abundance in the source, leading to initial melts enriched in components from low-melting-point phases. For instance, in the mantle, clinopyroxene and orthopyroxene contribute disproportionately to early melts compared to olivine, influencing the silica and alkali content of the resulting liquid.[7] The mineralogy of the source rock dictates the sequence and extent of melting, with low-melting minerals such as quartz and alkali feldspar in crustal sources or clinopyroxene and plagioclase in mantle sources melting preferentially at the solidus. In mafic sources like peridotite, partial melting at degrees of 10-20% typically produces basaltic compositions, as the olivine-pyroxene assemblage yields melts richer in MgO and CaO but lower in SiO₂ relative to the source. Conversely, felsic crustal sources, such as granitic gneisses, generate rhyolitic melts at lower fractions (often <10%), due to the dominance of quartz-feldspar assemblages that favor silica-enriched liquids. These differences arise because mafic minerals require higher temperatures to melt, resulting in more refractory residues in ultramafic sources compared to the leucocratic residues from felsic rocks.[1][9][10] Bulk chemistry further modulates the solidus temperature and melt fertility through variations in major element ratios, particularly Fe-Mg-Ca-Si. Higher Mg/Fe ratios in depleted peridotites elevate the solidus by stabilizing forsteritic olivine, shifting the onset of melting to higher temperatures and producing less fertile sources with lower melt fractions. In contrast, Ca- and Si-enriched compositions lower the solidus, enhancing melt productivity; for example, the addition of CaO promotes plagioclase instability, facilitating earlier melting in the assemblage. These shifts in the solidus curve directly impact the volume of melt generated, with Fe-rich sources exhibiting broader melting intervals due to lowered liquidus temperatures. Seminal experimental studies confirm that such bulk compositional variations can alter melt fractions by up to 10% for a given temperature.[11][12] Representative examples illustrate these controls: mid-ocean ridge basalts (MORB) derive from partial melting (5-20%) of depleted mantle peridotite, yielding tholeiitic compositions low in incompatible elements due to the refractory, olivine-rich residue. Ocean island basalts (OIB), however, stem from enriched mantle sources with higher Fe and volatiles, resulting in alkalic melts at lower degrees (<5%) and more evolved signatures from pyroxenite or recycled components within the peridotite. These distinctions highlight how source heterogeneity drives magmatic diversity without invoking external variables like pressure.[13][4][14]Temperature and Pressure Effects
Temperature plays a fundamental role in initiating partial melting by determining when a rock crosses its solidus temperature, the point at which the first melt forms. For dry mantle peridotite, this solidus typically ranges from approximately 1200°C to 1400°C, varying with bulk composition such as alkali content and Mg#; higher alkalis lower the solidus, while more magnesian compositions raise it slightly.[15] Once the temperature exceeds the solidus but remains below the liquidus (the temperature for complete melting), partial melting occurs, producing a melt fraction that increases with rising temperature.[15] Pressure exerts a stabilizing influence on the solid phases, elevating the solidus temperature and thereby suppressing melting at greater depths. In mantle peridotite, the solidus rises by about 100 °C per GPa near 1–2 GPa, reflecting the positive slope of the solidus in pressure-temperature (P-T) space; for instance, experimental fits indicate a nonlinear increase, with temperatures reaching around 1250°C at 1 GPa and 1350°C at 2 GPa for typical compositions.[15][16] This effect means that higher pressures favor the persistence of solid minerals over melt formation, and adiabatic decompression—where rising mantle material cools along a shallow adiabat—can effectively bring the rock's temperature above the lowering solidus, promoting melting onset.[15] In P-T phase diagrams, the solidus and liquidus appear as upward-sloping lines bounding the partial melting field, with the solidus marking the boundary between subsolidus assemblages and the onset of melt. A rock's P-T path, governed by tectonic processes, intersects the solidus to initiate partial melting; for example, if the path enters the melting interval without reaching the liquidus, only a portion of the rock melts, leaving a refractory residue.[15] These diagrams highlight how the width of the melting interval narrows at higher pressures due to the converging solidus and liquidus slopes.[16] Piston-cylinder experiments have confirmed pressure's role in stabilizing minerals like garnet over melt production, particularly in the 2–3 GPa range where garnet lherzolite assemblages dominate. At these pressures, the solidus cusp associated with the garnet-spinel transition raises the melting temperature by tens of degrees, as garnet's stability inhibits low-degree melting compared to spinel-feldspar fields at shallower depths.[16][17] Such experiments, conducted up to 3 GPa, demonstrate that increasing pressure shifts phase boundaries, favoring solid phases and delaying melt formation until higher temperatures are achieved.[17] Volatiles like water can enhance melting by further depressing the solidus, but the baseline dry effects underscore pressure's dominant stabilizing influence.[15]Role of Volatiles
Volatiles such as water (H₂O) and carbon dioxide (CO₂) play a pivotal role in partial melting by significantly lowering the solidus temperature of mantle rocks, thereby facilitating the generation of melt at conditions otherwise subsolidus. In mantle peridotite, the addition of H₂O depresses the solidus by approximately 250–300°C at 2–3 GPa for concentrations of 1–2 wt% in the melt, promoting the stability of hydrous phases like amphibole (e.g., pargasite) at pressures below 3 GPa. CO₂ exerts a milder effect, reducing the solidus by about 7°C per wt% at similar pressures, though its influence becomes more pronounced in the presence of H₂O due to enhanced solubility. These depressions arise from the incorporation of volatiles into silicate structures, weakening bonding and stabilizing liquid phases over solids.[18][19] The presence of volatiles not only initiates melting at lower temperatures but also increases the melt fraction for a given temperature excess above the solidus, enhancing overall melting efficiency. H₂O addition typically results in melts enriched in alumina (Al₂O₃) and depleted in magnesia (MgO) and lime (CaO), while maintaining relatively constant silica (SiO₂) content, leading to somewhat more siliceous compositions compared to anhydrous melts. CO₂, while less effective at solidus depression, expands the melt volume due to its lower density and further reduces melt viscosity—by up to two orders of magnitude at concentrations around 3.5 wt% near the glass transition—facilitating melt migration. Phase diagrams incorporating volatile isopleths illustrate these effects, showing curved solidus boundaries that shift downward with increasing H₂O or CO₂ content, and highlighting the expanded stability fields of fluxing phases like phlogopite in potassic systems.[18][19][20] In geological settings, slab-derived fluids rich in H₂O and CO₂ from subducting oceanic lithosphere trigger partial melting in the overlying mantle wedge, where dehydration reactions release volatiles at depths of 60–130 km. These fluids lower the local solidus, enabling hydrous melting at temperatures below 1000°C with H₂O contents exceeding 7 wt%, and producing arc magmas with elevated volatile signatures compared to mid-ocean ridge basalts. The fluxing action of these volatiles, particularly through amphibole and phlogopite breakdown, sustains melt production in subduction zones, influencing magma compositions toward more calc-alkaline varieties.[21][18]Melting Mechanisms
Decompression Melting
Decompression melting is a key process in mantle petrology where hot, solid mantle rock undergoes partial melting due to a reduction in pressure during adiabatic ascent, without the addition of external heat. As the mantle upwells, the decrease in pressure lowers the solidus temperature, causing the pressure-temperature (P-T) trajectory to intersect the solidus curve and initiate melting. This mechanism is passive and relies on the inherent thermal structure of the mantle, where the rock's temperature remains roughly constant or decreases slightly along the ascent path.[22][23] This process is most commonly associated with divergent tectonic boundaries, particularly beneath mid-ocean ridges, where convective upwelling of the asthenospheric mantle occurs to accommodate plate separation. Melting typically initiates at depths of 60–150 km and progresses to shallower levels, with melt fractions increasing progressively as pressure drops; fractions can reach up to 20% at depths less than 100 km under typical mantle conditions. The P-T path of the ascending mantle diverges from the dry adiabat because the latent heat of fusion is absorbed during melting, leading to a steeper, cooler trajectory compared to isentropic ascent. Faster ascent rates, such as those exceeding 100 mm/year, enhance melt productivity by reducing conductive heat loss to the surrounding lithosphere, allowing more extensive crossing of the solidus.[22][23][24] Decompression melting generates mid-ocean ridge basalts (MORB), which are derived from the aggregated partial melts of peridotite in the upwelling column and form the basaltic oceanic crust. Seismic observations provide evidence for this process, revealing low-velocity zones beneath mid-ocean ridges that are interpreted as regions containing 1–10% interconnected melt, confirming the presence of partial melting in the asthenosphere. Minor volatiles, such as water, can slightly lower the solidus and increase melt extents in these settings.[22][25][26]Flux Melting
Flux melting, also known as fluid-induced melting, occurs when water (H₂O) and carbon dioxide (CO₂) from subducting oceanic slabs or overlying sediments infiltrate the overlying mantle peridotite, dramatically lowering its solidus temperature and initiating partial melting to form hydrous magmas.[27][28] This process is driven by the release of volatiles during metamorphic dehydration of the subducting slab, which flux into the mantle wedge and react with peridotite minerals, such as olivine and pyroxenes, to produce low-degree hydrous melts enriched in incompatible elements.[29][30] A key aspect of flux melting involves a reaction series beginning with serpentinization of the subducting slab's ultramafic components at shallow depths, where hydration forms serpentine minerals under hydrous conditions, followed by progressive dehydration at greater depths that releases H₂O-rich fluids into the mantle.[31] These fluids then trigger melting in the mantle wedge peridotite through hydrous reactions, such as the breakdown of olivine to form amphibole or phlogopite, leading to the generation of small melt fractions (typically 1-5%).[32] The resulting melts migrate upward via porous flow through interconnected grain boundaries in the peridotite matrix, facilitated by the low viscosity of hydrous melts, which allows efficient extraction and ascent toward the surface.[29] This mechanism contrasts with baseline temperature-pressure conditions in the mantle, as the volatile influx enables melting at temperatures 200-300°C below the dry solidus.[33] Flux melting predominantly takes place at depths of 100-200 km in the mantle wedge of subduction zones, where slab-derived fluids interact with hot peridotite at temperatures around 1000-1200°C and pressures of 3-6 GPa, producing arc basalts characterized by distinctive trace element signatures such as enrichments in large-ion lithophile elements (e.g., Ba, Rb) and depletions in high-field-strength elements (e.g., Nb, Ta).[26][34] Experimental petrology demonstrates that even 1% H₂O can depress the solidus of peridotite by approximately 300°C, shifting the onset of melting to lower temperatures and promoting the formation of volatile-bearing magmas.[33][35] Prominent examples of flux melting are observed in Andean volcanism, particularly along the Central Volcanic Zone where subducting Nazca plate sediments and altered oceanic crust release volatiles that flux the mantle, generating andesitic to basaltic magmas responsible for volcanoes like Lascar and Nevado de Longaví.[36][37] These settings highlight how flux-induced partial melting contributes to the chemical diversity of arc magmas, with trace element patterns reflecting the influx of slab-derived components.[38]Heat Conduction Melting
Heat conduction melting represents a relatively slow mechanism of partial melting in the Earth's crust, where thermal energy diffuses from a hotter source into surrounding cooler rocks, gradually elevating temperatures until they exceed the solidus. This process primarily involves the transfer of heat from intruding hot mafic magmas, such as those derived from mantle sources, or from ascending mantle plumes that pond at the base of the lithosphere. In continental settings, underplating—where basaltic magma accumulates beneath the crust—facilitates conductive heating of overlying felsic or metasedimentary rocks, which have lower melting points, leading to localized partial melting without significant involvement of pressure changes or external fluids.[39][40] The conditions for heat conduction melting favor slower rates compared to decompression or flux mechanisms, often producing low melt fractions typically below 10%, as heat diffusion is inefficient over short timescales. This mechanism is prevalent in continental rift zones, where thinning crust allows mantle-derived heat to conduct upward, and in hotspot provinces influenced by mantle plumes, where the hot plume material establishes a steep thermal gradient at the lithosphere-asthenosphere boundary. The evolution of the thermal gradient is key: initial contact heating creates a localized high-temperature zone that propagates outward via conduction, potentially sustained over millions of years if the heat source persists, such as after lithospheric foundering or plume impingement. In such scenarios, radiogenic heat from the crust can amplify the effect, extending melting to mid-crustal levels.[41][40] Geophysical models of conductive heat flow illustrate this process in the formation of continental flood basalts, such as the Deccan Traps, where a mantle plume's thermal influence conducts heat into the overlying lithosphere and lower crust, generating initial melts that underplate and further promote crustal melting. These models demonstrate that sustained heat input from plumes can achieve sufficient temperatures for partial melting over protracted periods, with melt extraction facilitated once fractions exceed rheological thresholds (around 20-30% in some protoliths). While often operating alongside decompression in plume-rift hybrids, pure conduction dominates in stable continental interiors post-initial upwelling.[42][39]Geological Significance
Magma Generation and Composition
Partial melting is the primary mechanism for generating magma, where only a portion of the source rock melts, leaving behind a solid residue known as restite. During this process, incompatible elements—such as potassium (K), sodium (Na), rubidium (Rb), and certain rare earth elements (REE)—preferentially partition into the melt phase due to their low compatibility with the solid minerals (partition coefficients D << 1). This results in the magma becoming enriched in silica (SiO₂) and alkalis, while the restite becomes depleted in these components. For instance, in mantle-derived partial melts, the initial low-degree melts extract these elements efficiently, leading to higher concentrations in the liquid relative to the bulk source.[43] The compositional diversity of magmas arises directly from the source rock and the degree of partial melting. Partial melting of mantle peridotite typically produces basaltic magmas with low SiO₂ content (45-55 wt%), high levels of iron (Fe), magnesium (Mg), and calcium (Ca), and relatively low alkalis. In contrast, partial melting of crustal rocks, such as amphibolites or granites, yields more evolved compositions: andesitic magmas (55-65 wt% SiO₂, intermediate Fe, Mg, Ca, Na, K) or rhyolitic magmas (65-75 wt% SiO₂, low Fe, Mg, Ca, and high alkalis). Trace elements serve as provenance tracers; for example, enriched light REE patterns in continental magmas indicate low-degree melting of lithospheric sources, distinguishing them from depleted mantle-derived basalts.[44] A fundamental aspect of partial melting is the melt-restite relationship, which governs the evolution of major element ratios in the magma. The degree of melting (F, the fraction of melt produced) directly controls composition: higher F values (e.g., 20-30%) generate more mafic magmas closer to the bulk source composition, as less selective partitioning occurs, whereas low F (e.g., 1-10%) produces felsic, incompatible-element-enriched melts. This relationship is modeled by batch melting equations, where the melt concentration (C_l) of an element relative to the source (C_o) is given by C_l / C_o = 1 / [F + D(1 - F)] for incompatible elements (D ≈ 0), emphasizing enrichment at low F. The restite, enriched in compatible minerals like olivine and pyroxene, retains a more refractory, mafic character.[2][43] Isotopic studies provide robust evidence linking magma compositions to partial melting sources. Strontium-neodymium (Sr-Nd) isotope systematics, for example, reveal source signatures in arc and continental magmas; depleted mantle sources yield low ⁸⁷Sr/⁸⁶Sr (≈0.702-0.703) and high εNd (+5 to +10), while crustal partial melts show higher ⁸⁷Sr/⁸⁶Sr (0.71-0.75) and negative εNd (-10 to -15). In metasedimentary sources, such as those in the Himalayas, disequilibrium melting causes slight isotopic variations (e.g., elevated ⁸⁷Sr/⁸⁶Sr in melts due to muscovite breakdown), but overall trends confirm derivation from specific protoliths via partial melting. These tracers have been used to connect ocean island basalts to enriched mantle plumes and subduction-related andesites to hybrid crustal-mantle sources.[45]| Magma Type | Source | SiO₂ (wt%) | Key Characteristics | Example Setting |
|---|---|---|---|---|
| Basaltic | Mantle peridotite | 45-55 | High Fe, Mg, Ca; low alkalis; incompatible elements moderately enriched at low F | Mid-ocean ridges, hotspots[44] |
| Andesitic | Lower crust (amphibolite) | 55-65 | Intermediate Fe, Mg, Ca, Na, K; trace elements indicate mixed provenance | Subduction zones[44] |
| Rhyolitic | Upper crust (granite) | 65-75 | Low Fe, Mg, Ca; high SiO₂, alkalis; highly enriched in incompatibles | Continental arcs[44] |