4 Igneous Processes and Volcanoes

Learning Objectives

By the end of this chapter, students should be able to:

Igneous rock is formed when liquid rock freezes into a solid rock. This molten material is called magma when it is in the ground and lava when it is on the surface. Only the Earth’s outer core is liquid; the Earth’s mantle and crust is naturally solid. However, there are a few minor pockets of magma that form near the surface where geologic processes cause melting. It is this magma that becomes the source for volcanoes and igneous rocks. This chapter will describe the classification of igneous rocks, the unique processes that form magmas, types of volcanoes and volcanic processes, volcanic hazards, and igneous landforms. 

Stream of orange molten lava traveling down gently sloping hill of cooled black and gray volcanic rock.
Figure 4.1: Lava flow in Hawai’i.

Lava cools quickly on the surface of the earth and forms tiny microscopic crystals. These are known as fine-grained extrusive, or volcanic, igneous rocks. Extrusive rocks are often vesicular, filled with holes from escaping gas bubbles. Volcanism is the process in which lava is erupted. Depending on the properties of the lava that is erupted, the volcanism can be drastically different, from smooth and gentle to dangerous and explosive. This leads to different types of volcanoes and different volcanic hazards.

The granite formation rises against a backdrop of blue sky and gray mountains. Half Dome's distinct shape, with its sheer vertical face on one side and smooth, rounded surface on the other, is prominently displayed.
Figure 4.2: Half Dome, an intrusive igneous batholith in Yosemite National Park.

In contrast, magma that cools slowly below the earth’s surface forms larger crystals which can be seen with the naked eye. These are known as coarse-grained intrusive, or plutonic, igneous rocks. This relationship between cooling rates and grain sizes of the solidified minerals in igneous rocks is important for interpreting the rock’s geologic history.

4.1 Classification of Igneous Rocks

Igneous rocks are classified based on texture and composition. Texture describes the physical characteristics of the minerals, such as grain size. This relates to the cooling history of the molten magma from which it came. Composition refers to the rock’s specific mineralogy and chemical composition. Cooling history is also related to changes that can occur to the composition of igneous rocks.

4.1.1 Texture

Sample of granite that contains large, easily visible mineral grains of various colors including pink, white, gray, and black.
Figure 4.3: Granite is a classic coarse-grained (phaneritic) intrusive igneous rock. The different colors are unique minerals. The black colors are likely two or three different minerals.

If magma cools slowly, deep within the crust, the resulting rock is called intrusive or plutonic. The slow cooling process allows crystals to grow large, giving intrusive igneous rock a coarse-grained or phaneritic texture. The individual crystals in phaneritic texture are readily visible to the unaided eye.

Sample of basalt that is dull black without any clearly visible grains.
Figure 4.4: Basalt is a classic fine-grained extrusive igneous rock.

When lava is extruded onto the surface, or intruded into shallow fissures near the surface and cools, the resulting igneous rock is called extrusive or volcanic. Extrusive igneous rocks have a fine-grained or aphanitic texture, in which the grains are too small to see with the unaided eye. The fine-grained texture indicates the quickly cooling lava did not have time to grow large crystals. These tiny crystals can be viewed under a petrographic microscope. In some cases, extrusive lava cools so rapidly it does not develop crystals at all. This non-crystalline material is not classified as minerals, but as volcanic glass. This is a common component of volcanic ash and rocks like obsidian.

Dull grayish tan rock sample that contains large. easily visible tan mineral grains embedded within a finer-grained grayish tan groundmass.
Figure 4.5: Porphyritic texture.

Some igneous rocks have a mix of coarse-grained minerals surrounded by a matrix of fine-grained material in a texture called porphyritic. The large crystals are called phenocrysts and the fine-grained matrix is called the groundmass or matrix. Porphyritic texture indicates the magma body underwent a multi-stage cooling history, cooling slowly while deep under the surface and later rising to a shallower depth or the surface where it cooled more quickly.

Tan-to-black rock sample containing very large shiny crystals.
Figure 4.6: Pegmatitic texture.

Residual molten material expelled from igneous intrusions may form veins or masses containing very large crystals of minerals like feldspar, quartz, beryl, tourmaline, and mica. This texture, which indicates a very slow crystallization, is called pegmatitic. A rock that chiefly consists of pegmatitic texture is known as a pegmatite. To give an example of how large these crystals can get, transparent cleavage sheets of pegmatitic muscovite mica were used as windows during the Middle Ages.

A round-shaped sample of dull black lava rock full of small air holes.
Figure 4.7: Scoria.

All magmas contain gases dissolved in solution called volatiles. As the magma rises to the surface, the drop in pressure causes the dissolved volatiles to come bubbling out of solution, like the fizz in an opened bottle of soda. The gas bubbles become trapped in the solidifying lava to create a vesicular texture, with the holes specifically called vesicles. The type of volcanic rock with common vesicles is called scoria.

A round-shaped sample of dull white lava rock full of small air holes.
Figure 4.8: Pumice.

An extreme version of scoria occurs when volatile-rich lava is very quickly quenched and becomes a meringue-like froth of glass called pumice. Some pumice is so full of vesicles that the density of the rock drops low enough that it will float.

An angular chunk of shiny black volcanic glass that has a conchoidal fracture.
Figure 4.9: Obsidian (volcanic glass). Note conchoidal fracture.

Lava that cools extremely quickly may not form crystals at all, even microscopic ones. The resulting rock is called volcanic glass. Obsidian is a rock consisting of volcanic glass. Obsidian as a glassy rock shows an excellent example of conchoidal fracture similar to the mineral quartz (see chapter 3).

Dull tan rock made from volcanic ash that contains numerous various-sized angular fragments of other volcanic materials embedded, ranging from tan to black in color.
Figure 4.10: Welded tuff.

When volcanoes erupt explosively, vast amounts of lava, rock, ash, and gases are thrown into the atmosphere. The solid parts, called tephra, settle back to earth and cool into rocks with pyroclastic textures. Pyro, meaning fire, refers to the igneous source of the tephra and clastic refers to the rock fragments. Tephra fragments are named based on size—ash (<2 mm), lapilli (2-64 mm), and bombs or blocks (>64 mm). Pyroclastic texture is usually recognized by the chaotic mix of crystals, angular glass shards, and rock fragments. Rock formed from large deposits of tephra fragments is called tuff. If the fragments accumulate while still hot, the heat may deform the crystals and weld the mass together, forming a welded tuff.

4.1.2 Composition

Composition refers to a rock’s chemical and mineral make-up . For igneous rock, composition is divided into four groups: felsic, intermediate, mafic, and ultramafic. These groups refer to differing amounts of silica, iron, and magnesium found in the minerals that make up the rocks. It is important to realize these groups do not have sharp boundaries in nature, but rather lie on a continuous spectrum with many transitional compositions and names that refer to specific quantities of mineralsAs an example, granite is a commonly-used term, but has a very specific definition which includes exact quantities of minerals like feldspar and quartz. Rocks labeled as ‘granite‘ in laymen applications can be several other rocks, including syenite, tonalite, and monzonite. To avoid these complications, the following figure presents a simplified version of igneous rock nomenclature focusing on the four main groups, which is adequate for an introductory student.

Rectangular diagram with the following axis labels: on the left-hand vertical axis is "vol % of minerals" which increases from 0 at the bottom to 100 at the top; on the bottom horizontal axis is "wt% SiO2" which increases from 40 at the right-hand side to 70 at the left-hand side; and on the top horizontal axis are the following rock names: rhyolite and granite on the left, then dacite and granodiorite to the right of those, then andesite and diorite to the right of those, then basalt and gabbro to the right of those, and finally komatiite and peridotite to the far right. Inside the rectangular diagram are color-coded mineral names: tan orthoclase fills the upper left of the rectangle, then gray quartz is parallel to the right of orthoclase, then white plagioclase to the right of quartz, dark gray pyroxene to the right of that, and olive-green olivine filling the far right-hand side of the rectangle. At the lower left of the rectangle are cream-colored muscovite, dark brown biotite, and dark green amphibole.
Figure 4.11: Mineral composition of common igneous rocks. Percentage of minerals is shown on the vertical axis. Percentage of silica is shown on the horizontal axis. Rock names at the top include a continuous spectrum of compositions grading from one into another.

Felsic refers to a predominance of the light-colored (felsic) minerals feldspar and silica in the form of quartz. These light-colored minerals have more silica as a proportion of their overall chemical formula. Minor amounts of dark-colored (mafic) minerals like amphibole and biotite mica may be present as well. Felsic igneous rocks are rich in silica (in the 65-75% range, meaning the rock would be 65-75% weight percent SiO2) and poor in iron and magnesium.

Intermediate is a composition between felsic and mafic. It usually contains roughly-equal amounts of light and dark minerals, including light grains of plagioclase feldspar and dark minerals like amphibole. It is intermediate in silica in the 55-60% range.

Mafic refers to a abundance of ferromagnesian minerals (with magnesium and iron, chemical symbols Mg and Fe) plus plagioclase feldspar. It is mostly made of dark minerals like pyroxene and olivine, which are rich in iron and magnesium and relatively poor in silica. Mafic rocks are low in silica, in the 45-50% range.

Ultramafic refers to the extremely mafic rocks composed of mostly olivine and some pyroxene which have even more magnesium and iron and even less silica. These rocks are rare on the surface, but make up peridotite, the rock of the upper mantle. It is poor in silica, in the 40% or less range.

On the figure above, the top row has both plutonic and volcanic igneous rocks arranged in a continuous spectrum from felsic on the left to intermediate, mafic, and ultramafic toward the right. Rhyolite thus refers to the volcanic and felsic igneous rocks, and granite thus refer to intrusive and felsic igneous rocks. Andesite and diorite likewise refer to extrusive and intrusive intermediate rocks (with dacite and granodiorite applying to those rocks with composition between felsic and intermediate). Basalt and gabbro are the extrusive and intrusive names for mafic igneous rocks, and peridotite is ultramafic, with komatiite as the fine-grained extrusive equivalent. Komatiite is a rare rock because volcanic material that comes direct from the mantle is not common, although some examples can be found in ancient Archean rocks. Nature rarely has sharp boundaries and the classification and naming of rocks often imposes what appear to be sharp boundary names onto a continuous spectrum.

To identify a rock, match the texture (below) with the composition (right). Textures: phaneritic (coarse grained), aphanitic (fine grained), porphyritic (large crystals in a fine matrix), vesicular (bubbly or frothy), glassy, pyroclastic (fragmental). Compositions. Felsic (>10% quartz, >50% feldspar, <15% mafic materials). Intermediate (>10% quartz, >50% plagioclase, >10% orthoclase, <50% mafic minerals). Mafic (20-85% plagioclase, 15-50% pyroxene, >35% olivine). Ultramafic (olivine & pyroxene). Granite: phaneritic, felsic. Diorite: phaneritic, intermediate. Gabbro: phaneritic, mafic. Peridotite: phaneritic, ultramafic. Rhyolite: aphanitic, felsic. Andesite: aphanitic, intermediate. Basalt: aphanitic, mafic. Porphyritic rhyolite: porphyritic, felsic. Porphyritic andesite: porphyritic, intermediate. Porphyritic basalt: porphyritic, mafic. Pumice: vesicular, felsic. Scoria: vesicular, mafic. Obsidian: glassy, felsic. Rhyolitic tuff or volcanic breccia: pyroclastic, felsic. Andesitic tuff or volcanic breccia: pyroclastic, intermediate. Basaltic tuff or volcanic breccia: pyroclastic, mafic.
Figure 4.12: Igneous rock classification table with composition as vertical columns and texture as horizontal rows.
Aphanitic/phaneritic rock types
Felsic composition
Solid, opaque, square shaped, smooth sided rock with specs of pink, black, gray, and cream particles
Granite
Solid, opaque, square shaped, smooth sided rock made up of fine grained pink particles
Rhyolite
Granite is a course-crystalline felsic intrusive rock. The presence of quartz is a good indicator of granite. Granite commonly has large amounts of salmon pink potassium feldspar and white plagioclase crystals that have visible cleavage planes. Granite is a good approximation for the continental crust, both in density and composition. Rhyolite is a fine-crystalline felsic extrusive rock. Rhyolite is commonly pink and will often have glassy quartz phenocrysts. Because felsic lavas are less mobile, it is less common than granite. Examples of rhyolite include several lava flows in Yellowstone National Park and the altered rhyolite that makes up the Grand Canyon of the Yellowstone.
Intermediate composition
Solid, opaque, roughly rectangular, rough sided rock with specs of black and gray particles
Diorite
Solid, opaque rock with a rough triangle shape that is made of mostly fine grained cream particles with few larger gray and pink particles
Andesite
Diorite is a coarse-crystalline intermediate intrusive igneous rock. Diorite is identifiable by it’s Dalmatian-like appearance of black hornblende and biotite and white plagioclase feldspar. It is found in its namesake, the Andes Mountains as well as the Henry and Abajo mountains of Utah. Andesite is a fine crystalline intermediate extrusive rock. It is commonly grey and porphyritic. It can be found in the Andes Mountains and in some island arcs (see Chapter 2). It is the fine grained compositional equivalent of diorite.
Mafic composition
Solid, opaque, roughly triangular, fine-grained rock made of small brown and black particles
Gabbro
Solid, opaque, porous rock with smooth edges that is bluish gray
Vesicular basalt
Gabbro is a coarse-grained mafic igneous rock, made with mainly mafic minerals like pyroxene and only minor plagioclase. Because mafic lava is more mobile, it is less common than basalt. Gabbro is a major component of the lower oceanic crust. Basalt is a fine-grained mafic igneous rock. It is commonly vesicular and aphanitic. When porphyritic, it often has either olivine or plagioclase phenocrysts. Basalt is the main rock which is formed at mid-ocean ridges, and is therefore the most common rock on the Earth’s surface, making up the entirety of the ocean floor (except where covered by sediment).

Table 4.1: Aphanitic and phaneritic rock types with images.

4.1.3 Igneous Rock Bodies

Igneous rocks are common in the geologic record, but surprisingly, it is the intrusive rocks that are more common. Extrusive rocks, because of their small crystals and glass, are less durable. Plus, they are, by definition, exposed to the elements of erosion immediately. Intrusive rocks, forming underground with larger, stronger crystals, are more likely to last. Therefore, most landforms and rock groups that owe their origin to igneous rocks are intrusive bodies. A significant exception to this is active volcanoes, which are discussed in a later section on volcanism. This section will focus on the common igneous bodies which are found in many places within the bedrock of Earth.

Hilly landscape with dark red and gray hills made of flat-lying sedimentary layers, with a black igneous dike cutting perpendicular to the hill ridge that goes straight through the sedimentary layers.
Figure 4.13: Dike of olivine gabbro cuts across Baffin Island in the Canadian Arctic.

When magma intrudes into a weakness like a crack or fissure and solidifies, the resulting cross-cutting feature is called a dike (sometimes spelled dyke). Because of this, dikes are often vertical or at an angle relative to the pre-existing rock layers that they intersect. Dikes are therefore discordant intrusions, not following any layering that was present. Dikes are important to geologists, not only for the study of igneous rocks themselves but also for dating rock sequences and interpreting the geologic history of an area. The dike is younger than the rocks it cuts across and, as discussed in the chapter on Geologic Time (chapter 7), may be used to assign actual numeric ages to sedimentary sequences, which are notoriously difficult to age date. 

Cliff that shows slightly angled parallel tan, gray, and black sedimentary layers, with a tan, white, and black-colored igneous sill visible between and parallel to the sedimentary layers.
Figure 4.14: Igneous sill intruding between Paleozoic strata in Nova Scotia.

Sills are another type of intrusive structure. A sill is a concordant intrusion that runs parallel to the sedimentary layers in the country rock. They are formed when magma exploits a weakness between these layers, shouldering them apart and squeezing between them. As with dikes, sills are younger than the surrounding layers and may be radioactively dated to study the age of sedimentary strata.

White and light gray rock formations that stick out vertically from the surrounding hillside which is covered in green and brown trees.
Figure 4.15: Quartz monzonite in the Cretaceous of Montana, USA.

A magma chamber is a large underground reservoir of molten rock. The path of rising magma is called a diapir. The processes by which a diapir intrudes into the surrounding native or country rock are not well understood and are the subject of ongoing geological inquiry. For example, it is not known what happens to the pre-existing country rock as the diapir intrudes. One theory is the overriding rock gets shouldered aside, displaced by the increased volume of magma. Another is the native rock is melted and consumed into the rising magma or broken into pieces that settle into the magma, a process known as stoping. It has also been proposed that diapirs are not a real phenomenon, but just a series of dikes that blend into each other. The dikes may be intruding over millions of years, but since they may be made of similar material, they would be appearing to be formed at the same time. Regardless, when a diapir cools, it forms an mass of intrusive rock called a pluton. Plutons can have irregular shapes, but can often be somewhat round.

The granite formation rises against a backdrop of blue sky and gray mountains. Half Dome's distinct shape, with its sheer vertical face on one side and smooth, rounded surface on the other, is prominently displayed.
Figure 4.16: Half Dome in Yosemite National Park, California, is a part of the Sierra Nevada batholith which is mostly made of granite.

When many plutons merge together in an extensive single feature, it is called a batholith. Batholiths are found in the cores of many mountain ranges, including the granite formations of Yosemite National Park in the Sierra Nevada of California. They are typically more than 100 km2 in area, associated with subduction zones, and mostly felsic in composition. A stock is a type of pluton with less surface exposure than a batholith, and may represent a narrower neck of material emerging from the top of a batholith. Batholiths and stocks are discordant intrusions that cut across and through surrounding country rock.

Cross sectional diagram showing three flat-lying sedimentary layers colored orange, yellow, and brown stacked at the bottom of the diagram and three upward-arched sedimentary layers colored orange, yellow, and brown at the top of the diagram. A gray intrusion labeled "laccolith" fills the space between the arched sedimentary layers and the flat sedimentary layers with a nearly vertical thin gray intrusion labeled "dike" feeding it from below.
Figure 4.17: The Henry Mountains in Utah are interpreted to be a laccolith, exposed by erosion of the overlying layers.
High mountains covered in snow that have an overall broad, dome-shaped top. There are green trees and shrubs in the foreground.
Figure 4.18: Laccolith forms as a blister in between sedimentary strata.

Laccoliths are blister-like, concordant intrusions of magma that form between sedimentary layers. The Henry Mountains of Utah are a famous topographic landform formed by this process. Laccoliths bulge upwards; a similar downward-bulging intrusion is called a lopolith.

QR code
Complete this interactive activity to check your understanding.
Click on the plus signs on the illustration for descriptions of several igneous features.

If you are using an offline version of this text, access this interactive activity via the QR code.

 

QR code
Take this quiz to check your comprehension of this section.

If you are using an offline version of this text, access the quiz for section 4.1 via the QR code.

 

4.2 Bowen’s Reaction Series

A rectangular diagram with a vertical axis representing temperature increasing upward on the left. On the right-hand side of the diagram are two arrows: one points downward and is labeled "Increasing Silica, Na, Al, K" and the other points upward with the label "Increasing Fe, Mg & K." Inside the diagram are various mineral labels; starting from the bottom to top, they are Quartz, Muscovite, and K-Spar with a bracket labeling them "Residual Phases (Last to crystallize)" and above those are two branches: the discontinuous series branching toward the upper left and the continuous series branching toward the upper right. The discontinuous series contains Biotite, then Amphibole to the upper left of that, then Pyroxene to the upper left, and then Olivine to the upper left; the continuous series contains Oligoclase, then Anorthite to the upper left. On the right of the diagram are the labels Felsic, Intermediate, Mafic, and Ultramafic from bottom to top.
Figure 4.19: Bowen’s Reaction Series. Higher temperature minerals shown at top (olivine) and lower temperature minerals shown at bottom (quartz).
Light green, glassy crystal of olivine.
Figure 4.20: Olivine, the first mineral to crystallize in a melt.

Bowen’s Reaction Series describes the temperature at which minerals crystallize when cooled, or melt when heated. The low end of the temperature scale where all minerals crystallize into solid rock, is approximately 700°C (1292°F). The upper end of the range where all minerals exist in a molten state, is approximately 1,250°C (2,282°F). These numbers reference minerals that crystallize at standard sea-level pressure, 1 bar. The values will be different for minerals located deep below the Earth’s surface due to the increased pressure, which affects crystallization and melting temperatures. However, the order and relationships are maintained.

In the figure, the righthand column lists the four groups of igneous rock from top to bottom: ultramafic, mafic, intermediate, and felsic. The down-pointing arrow on the far right shows increasing amounts of silica, sodium, aluminum, and potassium as the mineral composition goes from ultramafic to felsic. The up-pointing arrow shows increasing ferromagnesian components, specifically iron, magnesium, and calcium. To the far left of the diagram is a temperature scale. Minerals near the top of diagram, such as olivine and anorthite (a type of plagioclase), crystallize at higher temperatures. Minerals near the bottom, such as quartz and muscovite, crystalize at lower temperatures.

Black and white photograph of Norman Bowen, a white man, in a suit with a black jacket and white collared shirt and white tie.
Figure 4.21: Norman L. Bowen.

The most important aspect of Bowen's Reaction Series is to notice the relationships between minerals and temperature. Norman L. Bowen (1887-1956) was an early 20th Century geologist who studied igneous rocks. He noticed that in igneous rocks, certain minerals always occur together and these mineral assemblages exclude other minerals. Curious as to why, and with the hypothesis in mind that it had to do with the temperature at which the rocks cooled, he set about conducting experiments on igneous rocks in the early 1900s. He conducted experiments on igneous rock—grinding combinations of rocks into powder, sealing the powders into metal capsules, heating them to various temperatures, and then cooling them.

Black and white photograph of Norman Bowen, a white man sitting in a chair, in a suit, holding and looking at an angular object, while a colleague, a white man seated behind him with glasses and a pipe and looking over his shoulder, also holds an angular object.
Figure 4.22: Norman L. Bowen and his colleague working at the Carnegie Institution of Washington Geophysical Laboratory.

When he opened the quenched capsules, he found a glass surrounding mineral crystals that he could identify under his petrographic microscope. The results of many of these experiments, conducted at different temperatures over a period of several years, showed that the common igneous minerals crystallize from magma at different temperatures. He also saw that minerals occur together in rocks with others that crystallize within similar temperature ranges, and never crystallize with other minerals. This relationship can explain the main difference between mafic and felsic igneous rocks. Mafic igneous rocks contain more mafic minerals, and therefore, crystallize at higher temperatures than felsic igneous rocks. This is even seen in lava flows, with felsic lavas erupting hundreds of degrees cooler than their mafic counterparts. Bowen’s work laid the foundation for understanding igneous petrology (the study of rocks) and resulted in his book, The Evolution of the Igneous Rocks in 1928.

QR code
Take this quiz to check your comprehension of this section.

If you are using an offline version of this text, access the quiz for section 4.2 via the QR code.

 

4.3 Magma Generation

Magma and lava contain three components: melt, solids, and volatiles. The melt is made of ions from minerals that have liquefied. The solids are made of crystallized minerals floating in the liquid melt. These may be minerals that have already cooled Volatiles are gaseous components—such as water vapor, carbon dioxide, sulfur, and chlorine—dissolved in the magma. The presence and amount of these three components affect the physical behavior of the magma and will be discussed more below.

4.3.1 Geothermal Gradient

Diagram showing temperature in kelvins increasing vertically downward with depth in km in the Earth. A red curve is drawn from the upper left corner of the diagram that shallowly dips down to the right until it reaches 410 km depth; this is labeled lithosphere. Below that, the curve steepens and goes deeper down until it shallows again around 2900 km depth. The part of the line from 410 to 660 km is labeled upper mantle and the line from 660 to 2900 km is labeled lower mantle. Below 2900 km, the curve steepens again down to 5150 km, and becomes vertical below 5150 km. From 2900 to 5150 km, the line is labeled outer core, and from 5150 to 6370 km, the line is labeled inner core.
Figure 4.23: Geothermal gradient.

Below the surface, the temperature of the Earth rises. This heat is caused by residual heat left from the formation of Earth and ongoing radioactive decay. The rate at which temperature increases with depth is called the geothermal gradient. The average geothermal gradient in the upper 100 km (62 mi) of the crust is about 25°C per kilometer of depth. So for every kilometer of depth, the temperature increases by about 25°C.

Diagram showing pressure increasing down on the vertical axis and temperature increasing to the right on the horizontal axis. There are two diagonal lines on the diagram: one is dark purple, labeled geotherm, that is a straight line from the upper left (0 on both axes) toward the lower right until it reaches about 125 km depth and 1250 degrees C temperature, where it then steepens to a nearly vertical line downward. The other line is light purple and is a slightly curved, steep line that goes from about 1100 degrees C and 0 km depth down steeply and slightly toward the right, labeled solidus.
Figure 4.24: Pressure-temperature diagram showing temperature in degrees Celsius on the x-axis and depth below the surface in kilometers (km) on the y-axis. The red line is the geothermal gradient and the green solidus line represents the temperature and pressure regime at which melting begins. Rocks at pressures and temperatures left of the green line are solid. If pressure/temperature conditions change so that rocks pass to the right of the green line, then they will start to melt.

The depth-temperature graph (see figure 4.23) illustrates the relationship between the geothermal gradient (geotherm, red line) and the start of rock melting (solidus, green line). The geothermal gradient changes with depth (which has a direct relationship to pressure) through the crust into upper mantle. The area to the left of the green line includes solid components; to the right is where liquid components start to form. The increasing temperature with depth makes the depth of about 125 kilometers (78 miles) where the natural geothermal gradient is closest to the solidus.

The temperature at 100 km (62 mi) deep is about 1,200°C (2,192°F). At bottom of the crust, 35 km (22 mi) deep, the pressure is about 10,000 bars. A bar is a measure of pressure, with 1 bar being normal atmospheric pressure at sea level. At these pressures and temperatures, the crust and mantle are solid. To a depth of 150 km (93 mi), the geothermal gradient line stays to the left of the solidus line. This relationship continues through the mantle to the coremantle boundary, at 2,880 km (1,790 mi).

The solidus line slopes to the right because the melting temperature of any substance depends on pressure. The higher pressure created at greater depth increases the temperature needed to melt rock. In another example, at sea level with an atmospheric pressure close to 1 bar, water boils at 100°C. But if the pressure is lowered, as shown on the video below, water boils at a much lower temperature.

QR code
Video 4.1: Boiling water at room temperature.

If you are using an offline version of this text, access this YouTube video via the QR code.

There are three principal ways rock behavior crosses to the right of the green solidus line to create molten magma: 1) decompression melting caused by lowering the pressure, 2) flux melting caused by adding volatiles (see more below), and 3) heat-induced melting caused by increasing the temperature. The Bowen’s Reaction Series shows that minerals melt at different temperatures. Since magma is a mixture of different minerals, the solidus boundary is more of a fuzzy zone rather than a well-defined line; some minerals are melted and some remain solid. This type of rock behavior is called partial melting and represents real-world magmas, which typically contain solid, liquid, and volatile components.

4.4 Volcanism

When magma emerges onto the Earth’s surface, the molten rock is called lava. A volcano is a type of land formation created when lava solidifies into rock. Volcanoes have been an important part of human society for centuries, though their understanding has greatly increased as our understanding of plate tectonics has made them less mysterious. This section describes volcano location, type, hazards, and monitoring.

4.4.1. Distribution and Tectonics

3D diagram showing a cross section into Earth's crust. On the left-hand side is a convergent plate boundary, with a plate of oceanic crust colliding into another plate of oceanic crust. The oceanic crust on the right-hand side subducts toward the left and a line of volcanoes forms on the overriding oceanic plate as a volcanic island arc. To the right of that, there is a shield volcano on the lithosphere beneath the ocean with a hot spot feeding the volcano from below. To the right of the shield volcano, there is a divergent plate boundary where two oceanic plates move away from each other. There is a transform plate boundary labeled to the left of the divergent boundary. To the right of the divergent boundary, oceanic crust moves toward the right, colliding with continental crust. The oceanic crust subducts beneath the continental crust, forming a chain of continental volcanoes on top of the continental crust. To the farthest right of the diagram, there is a continental rift zone where the continental crust is splitting apart.
Figure 4.25: Association of volcanoes with plate boundaries.

Most volcanoes are interplate volcanoes. Interplate volcanoes are located at active plate boundaries created by volcanism at mid-ocean ridges, subduction zones, and continental rifts. The prefix “inter-“ means between. Some volcanoes are intraplate volcanoes. The prefix “intra-“ means within, and intraplate volcanoes are located within tectonic plates, far removed from plate boundaries. Many intraplate volcanoes are formed by hotspots.

Volcanoes at Mid-ocean Ridges

World map with continents and oceans labeled. Multiple series of spreading ridges are drawn and labeled as red line segments on the map: major ones are found in an upside-down Y shape in the Indian Ocean, bordering the Pacific Ocean basin, running through the approximate center of the North and South Atlantic Ocean basins, and in the eastern corner of Africa. Small red dots mark the spots of recent subaerial volcanoes; they are found along the spreading ridge locations along with at hot spots and sparsely scattered on the Asian continent.
Figure 4.26: Map of spreading ridges throughout the world.

Most volcanism on Earth occurs on the ocean floor along mid-ocean ridges, a type of divergent plate boundary (see chapter 2). These interplate volcanoes are also the least observed and famous, since most of them are located under 3,000-4,500 m (10,000-15,000 ft) of ocean and the eruptions are slow, gentle, and oozing. One exception is the interplate volcanoes of Iceland. The diverging and thinning oceanic plates allow hot mantle rock to rise, releasing pressure and causing decompression melting. Ultramafic mantle rock, consisting largely of peridotite, partially melts and generates magma that is basaltic. Because of this, almost all volcanoes on the ocean floor are basaltic. In fact, most oceanic lithosphere is basaltic near the surface, with phaneritic gabbro and ultramafic peridotite underneath.

Underwater photo with a dark bluish hue of lumpy, bulbous chunks of black basalt on sea floor.
Figure 4.27: Pillow basalt on sea floor near Hawai’i.

When basaltic lava erupts underwater it emerges in small explosions and/or forms pillow-shaped structures called pillow basalts. These seafloor eruptions enable entire underwater ecosystems to thrive in the deep ocean around mid-ocean ridges. This ecosystem exists around tall vents emitting black, hot mineral-rich water called deep-sea hydrothermal vents, also known as black smokers.

A black smoker hydrothermal vent at the bottom of the sea floor. There is a plume of black smoke coming from a cone-shaped extrusion of rock and a colony of tube worms are attached to the cone-shaped rock.
Figure 4.28: Black smoker hydrothermal vent with a colony of giant (6’+) tube worms.
World map showing worldwide distribution of hydrothermal vent fields as red dots on the map; tectonic boundaries are drawn as thin black lines on the map. The hydrothermal vent fields are found along all of the mid-ocean ridges in the ocean basins.
Figure 4.29: Distribution of hydrothermal vent fields.

Without sunlight to support photosynthesis, these organisms instead utilize a process called chemosynthesis. Certain bacteria are able to turn hydrogen sulfide (H2S), a gas that smells like rotten eggs, into life-supporting nutrients and water. Larger organisms may eat these bacteria or absorb nutrients and water produced by bacteria living symbiotically inside their bodies. The videos show some of the ecosystems found around deep-sea hydrothermal vents.

QR code
Video 4.2: Updating the deep–diving submarine at 50-years-old.

If you are using an offline version of this text, access this YouTube video via the QR code.

QR code
Video 4.3: Incredible views on board the deep-sea vessel.

If you are using an offline version of this text, access this YouTube video via the QR code.

Volcanoes at Subduction Zones

World map showing with major tectonic plates labeled; plate boundaries are marked as black lines and active volcanoes are marked as red dots. Active volcanoes are shown along various plate boundaries, especially around the edts of the Pacific Ocean basin where they are labeled "Ring of Fire."
Figure 4.30: Distribution of volcanoes on the planet. Click here for an interactive map of volcano distributions.

The second most commonly found location for volcanism is adjacent to subduction zones, a type of convergent plate boundary (see chapter 2). The process of subduction expels water from hydrated minerals in the descending slab, which causes flux melting in the overlying mantle rock. Because subduction volcanism occurs in a volcanic arc, the thickened crust promotes partial melting and magma differentiation. These evolve the mafic magma from the mantle into more silica-rich magma. The Ring of Fire surrounding the Pacific Ocean, for example, is dominated by subduction-generated eruptions of mostly silica-rich lava; the volcanoes and plutons consist largely of intermediate-to-felsic rock such as andesite, rhyolite, pumice, and tuff.

Volcanoes at Continental Rifts

Barren landscape with gently rolling hills covered in black, brown, and red chunks of volcanic rock in the foreground, a large flat basin in the mid part of the photo, and a mountain in the distant background.
Figure 4.31: Basaltic cinder cones of the Black Rock Desert near Beaver, Utah.

Some volcanoes are created at continental rifts, where crustal thinning is caused by diverging lithospheric plates, such as the East African Rift Basin in Africa. Volcanism caused by crustal thinning without continental rifting is found in the Basin and Range Province in North America. In this location, volcanic activity is produced by rising magma that stretches the overlying crust. Lower crust or upper mantle material rises through the thinned crust, releases pressure, and undergoes decompression-induced partial melting. This magma is less dense than the surrounding rock and continues to rise through the crust to the surface, erupting as basaltic lava. These eruptions usually result in flood basalts, cinder cones, and basaltic lava flows (see video). Relatively young cinder cones of basaltic lava can be found in south-central Utah, in the Black Rock Desert Volcanic Field, which is part of the zone of Basin and Range crustal extension. These Utah cinder cones and lava flows started erupting around 6 million years ago, with the last eruption occurring 720 years ago.

QR code
Video 4.4: Basin and range volcanic processes.

If you are using an offline version of this text, access this YouTube video via the QR code.

Hotspots

Two cross sectional diagrams: the top diagram shows horizontal layers with a magma plume rising vertically through them and a volcano on top of the layers. The second diagram shows horizontal layers with the top layer moving toward the left; a magma plume rises vertically through them and a chain of volcanoes is formed on the top layer which is moving toward the left as the magma plume creates volcanoes on top.
Figure 4.32: Diagram showing a non-moving source of magma (mantle plume) and a moving overriding plate.

Hotspots are the main source of intraplate volcanism. Hotspots occur when lithospheric plates glide over a hot mantle plume, an ascending column of solid heated rock originating from deep within the mantle. The mantle plume generates melts as material rises, with the magma rising even more. When the ascending magma reaches the lithospheric crust, it spreads out into a mushroom-shaped head that is tens to hundreds of kilometers across.

Shaded relief map centered on Idaho, with small portions of the surrounding states shown too. The track of a hot spot is annotated on the map and color-coded at each past location. The hotspot started near the Idaho-Oregon-Nevada border with the label 16.1 which indicates it was there 16.1 million years ago, then moved relatively east-northeastward toward its present location near the Wyoming-Idaho-Montana border which is labeled 0.6-2.1 which indicates it has been there 2.1 to 0.6 million years ago. Note that the North American plate was moving over the hot spot, not that the hot spot was moving under North America.
Figure 4.33: The track of the Yellowstone hotspot, which shows the age of different eruptions in millions of years ago.

Since most mantle plumes are located beneath the oceanic lithosphere, the early stages of intraplate volcanism typically take place underwater. Over time, basaltic volcanoes may build up from the sea floor into islands, such as the Hawaiian Islands. Where a hotspot is found under a continental plate, contact with the hot mafic magma may cause the overlying felsic rock to melt and mix with the mafic material below, forming intermediate magma. Or the felsic magma may continue to rise, and cool into a granitic batholith or erupt as a felsic volcano. The Yellowstone caldera is an example of hotspot volcanism that resulted in an explosive eruption.

Map of the Hawaii-Emperor seamount chain and seafloor topography. The Aleutian trench and islands run approximately east-to-west along the top of the map, the Emperor seamount chain runs approximately north-to-south near the left-hand side of the map, and the Hawaiian seamount chain runs approximately west-northwest-to-east-southeast near the bottom of the map.
Figure 4.34: The Hawaiian–Emperor seamount and island chain.

A zone of actively erupting volcanism connected to a chain of extinct volcanoes indicates intraplate volcanism located over a hotspot. These volcanic chains are created by the overriding oceanic plate slowly moving over a hotspot mantle plume. These chains are seen on the seafloor and continents and include volcanoes that have been inactive for millions of years. The Hawaiian Islands on the Pacific Oceanic plate are the active end of a long volcanic chain that extends from the northwest Pacific Ocean to the Emperor Seamounts, all the way to the to the subduction zone beneath the Kamchatka Peninsula. The overriding North American continental plate moved across a mantle plume hotspot for several million years, creating a chain of volcanic calderas that extends from Southwestern Idaho to the presently active Yellowstone caldera in Wyoming.

The three-minute video (below) illustrates hotspot volcanoes.

QR code
Video 4.5: What is a volcanic hotspot?

If you are using an offline version of this text, access this YouTube video via the QR code.

4.4.2 Volcano Features and Types

There are several different types of volcanoes based on their shape, eruption style, magmatic composition, and other aspects.

QR code
Complete this interactive activity to check your understanding.

If you are using an offline version of this text, access this interactive activity via the QR code.

 

Large mountain topped with snow with an ash-covered cone volcano growing on the right flank of the mountain. In the foreground is a yellow-to-green field with rolling green hills toward the large mountain in the back.
Figure 4.35: Mt. Shasta in Washington state with Shastina, its parasitic cone.

The figure shows the main features of a typical stratovolcano: 1) magma chamber, 2) upper layers of lithosphere, 3) the conduit or narrow pipe through which the lava erupts, 4) the base or edge of the volcano, 5) a sill of magma between layers of the volcano, 6) a diapir or feeder tube to the sill, 7) layers of tephra (ash) from previous eruptions, 8 & 9) layers of lava erupting from the vent and flowing down the sides of the volcano, 10) the crater at the top of the volcano, 11) layers of lava and tephra on (12), a parasitic cone. A parasitic cone is a small volcano located on the flank of a larger volcano such as Shastina on Mount Shasta. Kilauea sitting on the flank of Mauna Loa is not considered a parasitic cone because it has its own separate magma chamber, 13) the vents of the parasite and the main volcano, 14) the rim of the crater, 15) clouds of ash blown into the sky by the eruption; this settles back onto the volcano and surrounding land.

Aerial view of the top of a large mountain that has a large hole at its top, filled with deep blue lake water. There is a small cone-shaped island in the lake visible in the lower right of the photo.
Figure 4.36: Oregon’s Crater Lake was formed about 7700 years ago after the eruption of Mount Mazama.

The largest craters are called calderas, such as the Crater Lake Caldera in Oregon. Many volcanic features are produced by viscosity, a basic property of a lava. Viscosity is the resistance to flowing by a fluid. Low viscosity magma flows easily more like syrup, the basaltic volcanism that occurs in Hawai’i on shield volcanoes. High viscosity means a thick and sticky magma, typically felsic or intermediate, that flows slowly, similar to toothpaste.

Shield Volcano

Brown-to black rubble-covered foreground with a broad mountaintop visible in the background.
Figure 4.37: Kilauea in Hawai’i.

The largest volcanoes are shield volcanoes. They are characterized by broad low-angle flanks, small vents at the top, and mafic magma chambers. The name comes from the side view, which resembles a medieval warrior’s shield. They are typically associated with hotspots, mid-ocean ridges, or continental rifts with rising upper mantle material. The low-angle flanks are built up slowly from numerous low-viscosity basaltic lava flows that spread out over long distances. The basaltic lava erupts effusively, meaning the eruptions are small, localized, and predictable.

Blocky lava flowing over a road in Hawaii. A tree is on fire on the left-hand side of the photo.
Figure 4.38: Eruption of Kilauea in 2018 produced high viscosity lava shown here crossing a road. This eruption caused much property damage.

Typically, shield volcano eruptions are not much of a hazard to human life—although non-explosive eruptions of Kilauea (Hawai’i) in 2018 produced uncharacteristically large lavas that damaged roads and structures. Mauna Loa (see USGS page) and Kilauea (see USGS page) in Hawai’i are examples of shield volcanoes. Shield volcanoes are also found in Iceland, the Galapagos Islands, Northern California, Oregon, and the East African Rift.

Satellite view of enormous tan-to-brown volcano. It has a large round shape and some small craters visible in its flank and caldera.
Figure 4.39: Olympus Mons, an enormous shield volcano on Mars, the largest volcano in the solar system, standing about two and a half times higher than Everest is above sea level.

The largest volcanic edifice in the Solar System is Olympus Mons on Mars. This (possibly extinct) shield volcano covers an area the size of the state of Arizona. This may indicate the volcano erupted over a hotspot for millions of years, which means Mars had little, if any, plate tectonic activity.

Grayish black outcrop of volcanic rock with undulating, ropey appearance.
Figure 4.40: Ropey pahoehoe lava.

Basaltic lava forms special landforms based on magma temperature, composition, and content of dissolved gases and water vapor. The two main types of basaltic volcanic rock have Hawaiian names—pahoehoe and aa. Pahoehoe might come from low-viscosity lava that flows easily into ropey strands.

Jagged, sharp-looking black volcanic rock with red hot lava showing underneath parts of the black rocks. There are three mountain peaks in the background.
Figure 4.41: Blocky a’a lava.

Aa (sometimes spelled a’a or ʻaʻā and pronounced “ah-ah”) is more viscous and has a crumbly blocky appearance. The exact details of what forms the two types of flows are still up for debate. Felsic lavas have lower temperatures and more silica, and thus are higher viscosity. These also form aa-style flows.

Orange-yellow molten lava sputtering upward from black volcanic rock. There is a gently-sloping lava flow traveling outward from the eruption.
Figure 4.42: Volcanic fissure and flow, which could eventually form a lava tube.

Low-viscosity, fast-flowing basaltic lava tends to harden on the outside into a tube and continue to flow internally. Once lava flow subsides, the empty outer shell may be left as a lava tube. Lava tubes, with or without collapsed roofs, make famous caves in Hawai’i, Northern California, the Columbia River Basalt Plateau of Washington and Oregon, El Malpais National Monument in New Mexico, and Craters of the Moon National Monument in Idaho.

Fissures are cracks that commonly originate from shield-style eruptions. Lava emerging from fissures is typically mafic and very fluid. The 2018 Kiluaea eruption included fissures associated with the lava flows. Some fissures are caused by the volcanic seismic activity rather than lava flows. Some fissures are influenced by plate tectonics, such as the common fissures located parallel to the divergent boundary in Iceland.

High tower of rock composed of vertical polygonal columns that are tan in color. There are green trees surrounding the base of the tower.
Figure 4.43: Devils Tower in Wyoming has columnar jointing.
Landscape composed of interlocking hexagonal tan, red, and brown basalt columns that vary in height.
Figure 4.44: Columnar jointing on Giant’s Causeway in Ireland.

Cooling lava can contract into columns with semi-hexagonal cross sections called columnar jointing. This feature forms the famous Devils Tower in Wyoming, possibly an ancient volcanic vent from which the surrounding layers of lava and ash have been removed by erosion. Another well-known exposed example of columnar jointing is the Giant’s Causeway in Ireland.

Stratovolcano

View of city buildings with towering snow-capped mountain in the background.
Figure 4.45: Mount Rainier towers over Tacoma, Washington.

A stratovolcano, also called a composite cone volcano, has steep flanks, a symmetrical cone shape, distinct crater, and rises prominently above the surrounding landscape. The term composite refers to the alternating layers of pyroclastic fragments like ash and bombs, and solidified lava flows of varying composition. Examples include Mount Rainier in Washington state and Mount Fuji in Japan.

Tall, nearly-symmetrical snow-capped mountain.
Figure 4.46: Mt. Fuji in Japan, a typical stratovolcano, symmetrical, increasing slope, visible crater at the top.

Stratovolcanoes usually have felsic to intermediate magma chambers, but can even produce mafic lavas. Stratovolcanoes have viscous lava flows and domes, punctuated by explosive eruptions. This produces volcanoes with steep flanks.

Lava Domes

Aerial view of the top of a large barren mountain that has a large hole at its top. There is a small bulbous shape growing in the hole with steam coming out of it.
Figure 4.47: Lava domes have started the rebuilding process at Mount St. Helens, Washington.

Lava domes are accumulations of silica-rich volcanic rock, such as rhyolite and obsidian. Too viscous to flow easily, the felsic lava tends to pile up near the vent in blocky masses. Lava domes often form in a vent within the collapsed crater of a stratovolcano, and grow by internal expansion. As the dome expands, the outer surface cools, hardens, and shatters, and spills loose fragments down the sides. Mount Saint Helens has a good example of a lava dome inside of a collapsed stratovolcano crater. Examples of stand-alone lava domes are Chaiten in Chile and Mammoth Mountain in California.

Caldera

Four cross sectional diagrams. The first diagram shows a blob of magma at the bottom with two narrow vents rising vertically from the blob, one near the center of the diagram and the other to the right of that one; at the surface is a layered volcano with ash and pumice erupting where the side vent reaches the surface. The second diagram shows the layered volcano has collapsed at its center with arrows pointing downward to show the collapse. The third diagram shows small amounts of steam rising from the surface where the layered volcano has collapsed. The fourth diagram is labeled "Today" and shows the modern-day volcano with Garfield Peak on the left side of the caldera and Llao Rock on the right side of the caldera, with a lake and Wizard Island in the center of the caldera.
Figure 4.48: Timeline of events at Mount Mazama.
View through a clearing in green trees of a lake with a cone-shaped island in it.
Figure 4.49: Wizard Island sits in the caldera at Crater Lake.

Calderas are steep-walled, basin-shaped depressions formed by the collapse of a volcanic edifice into an empty magma chamber. Calderas are generally very large, with diameters of up to 25 km (15.5 mi). The term caldera specifically refers to a volcanic vent; however, it is frequently used to describe a volcano type. Caldera volcanoes are typically formed by eruptions of high-viscosity felsic lava having high volatiles content.

Crater Lake, Yellowstone, and the Long Valley Caldera are good examples of this type of volcanism. The caldera at Crater Lake National Park in Oregon was created about 6,800 years ago when Mount Mazama, a composite volcano, erupted in a huge explosive blast. The volcano ejected large amounts of volcanic ash and rapidly drained the magma chamber, causing the top to collapse into a large depression that later filled with water. Wizard Island in the middle of the lake is a later resurgent lava dome that formed within the caldera basin.

Shaded relief map centered on Wyoming, with small portions of the surrounding states shown too. The caldera and related volcanic rocks, along with faults and earthquake epicenters, are annotated on the map and color-coded. The Yellowstone Caldera is outlined in purple, located in the west-central portion of Wyoming. Boundaries of older calderas are outlined in green with a dashed line where uncertain; there's an older caldera near the Idaho-Wyoming border, located in Idaho. There's an uncertain caldera from just west of the Wyoming border into Wyoming, covering the west-central portion of the state. There are numerous earthquake epicenters scattered throughout the Yellowstone caldera and also clustered just northwest of the caldera.
Figure 4.50: Map of calderas and related rocks around Yellowstone.

The Yellowstone volcanic system erupted three times in the recent geologic past—2.1, 1.3, and 0.64 million years ago—leaving behind three caldera basins. Each eruption created large rhyolite lava flows as well as pyroclastic flows that solidified into tuff formations. These extra-large eruptions rapidly emptied the magma chamber, causing the roof to collapse and form a caldera. The youngest of the three calderas contains most of Yellowstone National Park, as well as two resurgent lava domes. The calderas are difficult to see today due to the amount of time since their eruptions and subsequent erosion and glaciation.

Yellowstone volcanism started about 17-million years ago as a hotspot under the North American lithospheric plate near the Oregon/Nevada border. As the plate moved to the southwest over the stationary hotspot, it left behind a track of past volcanic activities. Idaho’s Snake River Plain was created from volcanism that produced a series of calderas and lava flows. The plate eventually arrived at its current location in northwestern Wyoming, where hotspot volcanism formed the Yellowstone calderas.

Map of the United States with state borders outlined. Prominent ash beds are outlined and color-coded, including three Yellowstone eruptions shaded pink. One of the pink outlines is labeled Mesa Falls ash bed and encircles most of the states of Wyoming, Colorado, Kansas, and Nebraska, and partially encircles the states Montana, South Dakota, Oklahoma, and Texas. Another pink outline is labeled Huckleberry Ridge ash bed and encircles the a large western portion of the United States. The third pink outline is labeled Lava Creek ash bed and encircles most of the western half of the United States. There is also a brown dashed outline labeled "Bishop ash bed" which encircles the entire southwest portion of the United States. There is a yellow elongated outline labeled "Mount St. Helens ash 1980" which covers a east-west-trending portion of southern Washington state.
Figure 4.51: Several prominent ash beds found in North America, including three Yellowstone eruptions shaded pink (Mesa Falls, Huckleberry Ridge, and Lava Creek), the Bisho Tuff ash bed (brown dashed line), and the modern May 18th, 1980 ash fall (yellow).

The Long Valley Caldera near Mammoth, California, is the result of a large volcanic eruption that occurred 760,000 years ago. The explosive eruption dumped enormous amounts of ash across the United States, in a manner similar to the Yellowstone eruptions. The Bishop Tuff deposit near Bishop, California, is made of ash from this eruption. The current caldera basin is 17 km by 32 km (10 mi by 20 mi), large enough to contain the town of Mammoth Lakes, major ski resort, airport, major highway, resurgent dome, and several hot springs.

Cinder Cone

Aerial view of brown cinder cone volcano with large opening at top; the surrounding landscape is low, shrubby green vegetation.
Figure 4.52: Sunset Crater, Arizona is a cinder cone.

Cinder cones are small volcanoes with steep sides, and made of pyroclastic fragments that have been ejected from a pronounced central vent. The small fragments are called cinders and the largest are volcanic bombs. The eruptions are usually short-lived events, typically consisting of mafic lavas with a high content of volatiles. Hot lava is ejected into the air, cooling and solidifying into fragments that accumulate on the flank of the volcano. Cinder cones are found throughout western North America.

Slightly grainy color photograph of a cinder cone volcano erupting an ash plume with a person wearing a brown plaid sombrero and red handkerchief looking on.
Figure 4.53: Soon after the birth of Parícutin in 1943.
Grayish tan stone church tower and building surrounded by black and brown blocky volcanic rocks.
Figure 4.54: Lava from Parícutin covered the local church and destroyed the town of San Juan, Mexico.

A recent and striking example of a cinder cone is the eruption near the village of Parícutin, Mexico that started in 1943. The cinder cone started explosively shooting cinders out of the vent in the middle of a farmer’s field. The volcanism quickly built up the cone to a height of over 90 m (300 ft) within a week, and 365 m (1,200 ft) within the first 8 months. After the initial explosive eruption of gases and cinders, basaltic lava poured out from the base of the cone. This is a common order of events for cinder cones: violent eruption, cone and crater formation, low-viscosity lava flow from the base. The cinder cone is not strong enough to support a column of lava rising to the top of the crater, so the lava breaks through and emerges near the bottom of the volcano. During nine years of eruption activity, the ashfall covered about 260 km2 (100 mi2) and destroyed the nearby town of San Juan.

Flood Basalts

World map of flood basalts. Includes Sierra Madre occidental, Columbia River basalt group (2), High Arctic LIP (3), North Atlantic LIP (3), Siberian Traps, Emeishan Traps, CAMP (3), Parana-Etendeka LIP (4), Afro-Arabian LIP (2), Karoo Ferrar LIP, Deccan traps, Rajmahal LIP, Gawler Range LIP, Whitsunday LIP (2). Note the largest is the Siberian Traps
Figure 4.55: World map of flood basalts. Note the largest is the Siberian Traps.

A rare volcanic eruption type, unobserved in modern times, is the flood basalt. Flood basalts are some of the largest and lowest viscosity types of eruptions known. They are not known from any eruption in human history, so the exact mechanisms of eruption are still mysterious. Some famous examples include the Columbia River Flood Basalts in Washington, Oregon, and Idaho, the Deccan Traps, which cover about 1/3 of the country of India, and the Siberian Traps, which may have been involved in the Earth’s largest mass extinction (see chapter 8).

Table of igneous rocks and related volcano types. Horizontal axis is arranged from low to high silica content (i.e. from ultramafic to felsic). First row shows the extrusive (surface) igneous rocks basalt, andesite, and rhyolite. Second row shows volcano types: mid-ocean ridge, shield, cinder cone, and strato (composite). Third row shows examples of each volcano: mid-atlantic ridge, Mauna Kea (Hawaii), Paricutin, and Mt. St. Helens. Fourth row shows intrusive rocks from mafic to felsic: Dunite, gabbro, diorige, granite. Fifth row shows common plate-tectonic settings: divergent oceanic hot spot, and convergent boundaries. Sixth row is typical composition: ultramafic, mafic, intermediate, and felsic.
Figure 4.56: Igneous rock types and related volcano types. Mid-ocean ridges and shield volcanoes represent more mafic compositions, and strato (composite) volcanoes generally represent a more intermediate or felsic composition and a convergent plate tectonic boundary. Note that there are exceptions to this generalized layout of volcano types and igneous rock composition.

4.4.3 Volcanic Hazards and Monitoring

Block diagram of a volcano with various hazards labeled: along the flanks of the volcano are a lava flow, lahar, two pyroclastic flows, and a landslide; erupting from the volcano are an eruption column and tephra with an eruption cloud, gas, and acid rain coming from the eruption column. There is a vent of magma feeding the volcano from below and a small lava dome on part of the side of the volcano.
Figure 4.57: General diagram of volcanic hazards.

While the most obvious volcanic hazard is lava, the dangers posed by volcanoes go far beyond lava flows. For example, on May 18, 1980, Mount Saint Helens (Washington, United States) erupted with an explosion and landslide that removed the upper 400 m (1,300 ft) of the mountain. The initial explosion was immediately followed by a lateral blast, which produced a pyroclastic flow that covered nearly 600 km2 (230 mi2) of forest with hot ash and debris. The pyroclastic flow moved at speeds of 80-130 kph (50-80 mph), flattening trees and ejecting clouds of ash into the air. The USGS video provides an account of this explosive eruption that killed 57 people.

QR code
Video 4.6: Mout St. Helens.

If you are using an offline version of this text, access this YouTube video via the QR code.

The body is laying down, encased in a cast formed from grayish white ash.
Figure 4.58: Human remains from the 79 CE eruption of Vesuvius.

In 79 AD, Mount Vesuvius, located near Naples, Italy, violently erupted sending a pyroclastic flow over the Roman countryside, including the cities of Herculaneum and Pompeii. The buried towns were discovered in an archeological expedition in the 18th century. Pompeii famously contains the remains (casts) of people suffocated by ash and covered by 10 feet (3 m) of ash, pumice lapilli, and collapsed roofs.

Two photographs of a large volcano. The left photo is labeled "Before," showing a large, conical volcano with snow on its flanks and a forest at its base. The right photo is labeled "After," showing the peak of the mountain has collapsed into a caldera and the landscape is barren.
Figure 4.59: Mount St. Helens, the day before the May 18th, 1980 eruption (left) and 4 months after the major eruption (right).
Black and white photo of a large volcano erupting with billowing smoke and dark flows down its flanks.
Figure 4.60: Image from the May 18, 1980, eruption of Mt. Saint Helens, Washington.

Pyroclastic Flows

Erupting volcano with billowing smoke erupting from its peak, traveling upward as well as laterally down the right flank of the volcano.
Figure 4.61: The material coming down from the eruption column is a pyroclastic flow.

The most dangerous volcanic hazard are pyroclastic flows (video). These flows are a mix of lava blocks, pumice, ash, and hot gases between 200°C-700°C (400°F-1,300°F). The turbulent cloud of ash and gas races down the steep flanks at high speeds up to 193 kph (120 mph) into the valleys around composite volcanoes. Most explosive, silica-rich, high viscosity magma volcanoes such as composite cones usually have pyroclastic flows. The rock tuff and welded tuff is often formed from these pyroclastic flows.

Black and white photo of a man sitting on a ridge, overlooking a destroyed city.
Figure 4.62: The remains of St. Pierre.

There are numerous examples of deadly pyroclastic flows. In 2014, the Mount Ontake pyroclastic flow in Japan killed 47 people. The flow was caused by magma heating groundwater into steam, which then rapidly ejected with ash and volcanic bombs. Some were killed by inhalation of toxic gases and hot ash, while others were struck by volcanic bombs. Two short videos below document eye-witness video of pyroclastic flows. In the early 1990s, Mount Unzen erupted several times with pyroclastic flows. The pyroclastic flow shown in this famous short video killed 41 people. In 1902, on the Caribbean Island Martinique, Mount Pelee erupted with a violent pyroclastic flow that destroyed the entire town of St. Pierre and killing 28,000 people in moments.

QR code
Video 4.7: Dome collapse and pyroclastic flow at Unzen Volcano.

If you are using an offline version of this text, access this YouTube video via the QR code.

Landslides and Landslide–Generated Tsunamis

Series of four schematic diagrams of Mt. St. Helens in a diagonal line from upper left to lower right. Starting at the upper left, the diagram shows a volcano with three domes labeled: Summit dome at the top, Cryptodome in the center base of the summit, and Goat Rocks dome along the right flank. The second diagram shows a landslide that collapsed the right flank of the volcano; the third diagram shows the initial explosions that resulted, erupting upward and toward the right, and the fourth diagram shows a vertical eruption column coming from the now-collapsed right flank of the volcano. On the lower left of the figure is two black and white photos of the eruption column. On the upper right of the figure is the following text: At 8:32 a.m., May 18, 1980, a 5.1 earthquake shook loose the north flank of Mount St. Helens, resulting in the largest known landslide in historic time. Removal of more than half a cubic mile of material released pressure and triggered a devastating lateral blast and ash-laden eruptive column.
Figure 4.63: Sequence of events for Mount St. Helens, May 18th, 1980. Note that an earthquake caused a landslide, which caused the “uncorking” of the mountain and started the eruption.

The steep and unstable flanks of a volcano can lead to slope failure and dangerous landslides. These landslides can be triggered by magma movement, explosive eruptions, large earthquakes, and/or heavy rainfall. During the 1980 Mount St. Helens eruption, the entire north flank of the volcano collapsed and released a huge landslide that moved at speeds of 160-290 kph (100-180 mph).

If enough landslide material reaches the ocean, it may cause a tsunami. In 1792, a landslide caused by the Mount Unzen eruption reached the Ariaka Sea, generating a tsunami that killed 15,000 people (see USGS page). When Mount Krakatau in Indonesia erupted in 1883, it generated ocean waves that towered 40 m (131 ft) above sea level. The tsunami killed 36,000 people and destroyed 165 villages.

Tephra

A man wearing a surgical mask sweeping white ash off the street. He is also wearing a conical hat, a red jacket, and brown pants. There is abundant white ash covering the street around and behind him.
Figure 4.64: Aman sweeps ash from an eruption of Kelud, Indonesia.

Volcanoes, especially composite volcanoes, eject large amounts of tephra (ejected rock materials), most notably ash (tephra fragments less than 0.08 inches [2 mm]). Larger tephra is heavier and falls closer to the vent. Larger blocks and bombs pose hazards to those close to the eruption such as at the 2014 Mount Ontake disaster in Japan discussed earlier.

Grayscale micrograph of volcanic ash particle with abundant pore space throughout the particle. A scale bar labeled 0 to 30 micrometers is in the upper right of the image.
Figure 4.65: Micrograph of silica particle in volcanic ash. A cloud of these is capable of destroying an aircraft or automobile engine.

Hot ash poses an immediate danger to people, animals, plants, machines, roads, and buildings located close to the eruption. Ash is fine grained (< 2mm) and can travel airborne long distances away from the eruption site. Heavy accumulations of ash can cause buildings to collapse. In people, it may cause respiratory issues like silicosis. Ash is destructive to aircraft and automobile engines, which can disrupt transportation and shipping services. In 2010, the Eyjafjallajökull volcano in Iceland emitted a large ash cloud into the upper atmosphere, causing the largest air-travel disruption in northern Europe since World War II. No one was injured, but the service disruption was estimated to have cost the world economy billions of dollars.

Volcanic Gases

As magma rises to the surface the confining pressure decreases, and allows dissolved gases to escape into the atmosphere. Even volcanoes that are not actively erupting may emit hazardous gases, such as carbon dioxide (CO2), sulfur dioxide (SO2), hydrogen sulfide (H2S), and hydrogen halides (HF, HCl, or HBr).

Carbon dioxide tends to sink and accumulate in depressions and basins. In volcanic areas known to emit carbon dioxide, low-lying areas may trap hazardous concentrations of this colorless and odorless gas. The Mammoth Mountain Ski Resort in California, is located within the Long Valley Caldera, is one such area of carbon dioxide-producing volcanism. In 2006, three ski patrol members died of suffocation caused by carbon dioxide after falling into a snow depression near a fumarole (info).

In rare cases, volcanism may create a sudden emission of gases without warning. Limnic eruptions (limne is Greek for lake), occur in crater lakes associated with active volcanism. The water in these lakes is supercharged with high concentrations of dissolved gases. If the water is physically jolted by a landslide or earthquake, it may trigger an immediate and massive release of gases out of solution. An analogous example would be what happens to vigorously shaken bottle of carbonated soda when the cap is opened. An infamous limnic eruption occurred in 1986 at Lake Nyos, Cameroon. Almost 2,000 people were killed by a massive release of carbon dioxide.

Lahars

Barren trees with a noticeable horizontal line throughout them at the same height, below which the trees are grayish tan from being covered in mud and above which the brown bark of the trees is visible.
Figure 4.66: Mud line shows the extent of lahars around Mount St. Helens.

Lahar is an Indonesian word and is used to describe a volcanic mudflow that forms from rapidly melting snow or glaciers. Lahars are slurries resembling wet concrete, and consist of water, ash, rock fragments, and other debris. These mudflows flow down the flanks of volcanoes or mountains covered with freshly-erupted ash and on steep slopes can reach speeds of up to 80 kph (50 mph).

Shaded relief map with Tacoma, Washington, at the upper left of the map and Mount Rainier at the lower right of the map. Three lahars are colored on the map: the Osceola Lahar in yellow that goes from Mount Rainier to Tacoma, generally following the path of the White River; the National Lahar in pink that goes from the south side of Mount Rainier toward the left, generally following the path of the Nisqually River; and the Electron Lahar in lilac that goes from Mount Rainier to Puyallup, generally following the Puyallup River.
Figure 4.67: Old lahars around Tacoma, Washington.

Several major cities, including Tacoma, are located on prehistoric lahar flows that extend for many kilometers across the flood plains surrounding Mount Rainier in Washington (see map). A map of Mount Baker in Oregon shows a similar potential hazard for lahar flows (see map). A tragic scenario played out recently, in 1985, when a lahar from the Nevado del Ruiz volcano in Colombia buried the town of Armero and killed an estimated 23,000 people.

Monitoring

Geologists use various instruments to detect changes or indications that an eruption is imminent. The three videos show different types of volcanic monitoring used to predict eruptions 1) earthquake activity; 2) increases in gas emission; and 3) changes in land surface orientation and elevation.

One video shows how monitoring earthquake frequency, especially special vibrational earthquakes called harmonic tremors, can detect magma movement and possible eruption. Another video shows how gas monitoring may be used to predict an eruption. A rapid increase of gas emission may indicate magma that is actively rising to surface and releasing dissolved gases out of solution, and that an eruption is imminent. The last video shows how a GPS unit and tiltmeter can detect land surface changes, indicating the magma is moving underneath it.

QR code
Video 4.8: Earthquake signals.

If you are using an offline version of this text, access this YouTube video via the QR code.

QR code
Video 4.9: Measuring gas emissions.

If you are using an offline version of this text, access this YouTube video via the QR code.

QR code
Video 4.10: Using tiltmeters and GPS to monitor a volcano.

If you are using an offline version of this text, access this YouTube video via the QR code.

QR code
Take this quiz to check your comprehension of this section.

If you are using an offline version of this text, access the quiz for section 4.4 via the QR code.

 

Summary

Igneous rock is divided into two major groups: intrusive rock that solidifies from underground magma, and extrusive rock formed from lava that erupts and cools on the surface. Magma is generated from mantle material at several plate tectonics situations by three types of melting: decompression melting, flux melting, or heat-induced melting. Magma composition is determined by differences in the melting temperatures of the mineral components (Bowen’s Reaction Series). The processes affecting magma composition include partial melting, magmatic differentiation, assimilation, and collision. Volcanoes come in a wide variety of shapes and sizes, and are classified by a multiple factors, including magma composition, and plate tectonic activity. Because volcanism presents serious hazards to human civilization, geologists carefully monitor volcanic activity to mitigate or avoid the dangers it presents.

QR code
Take this quiz to check your comprehension of this chapter.

If you are using an offline version of this text, access the quiz for chapter 4 via the QR code.

 

URLs Listed Within This Chapter

USGS page of Mauna Loa: https://www.usgs.gov/volcanoes/mauna-loa 

USGS page on Kilauea: https://www.usgs.gov/volcanoes/kilauea

Pyroclastic flows video: https://volcanoes.usgs.gov/vsc/movies/movie_101/PF_Animation.mp4

USGS page on volcano landslides that trigger waves and tsunamis: https://volcanoes.usgs.gov/Imgs/Jpg/Unzen/MayuyamaSlide_caption.html

Ski patrol’s fatal fall into a volcanic fumarole: https://pubmed.ncbi.nlm.nih.gov/19364170/

Text References

  1. Arndt, N.T., 1994, Chapter 1 Archean Komatiites, in K.C. Condie, editor, Developments in Precambrian Geology: Elsevier, p. 11–44.
  2. Bateman, P.C., and Chappell, B.W., 1979, Crystallization, fractionation, and solidification of the Tuolumne Intrusive Series, Yosemite National Park, California: Geological Society of America Bulletin, v. 90, no. 5, p. 465–482., https://doi.org/10.1130/0016-7606(1979)90<465:CFASOT>2.0.CO;2
  3. Boehler, R., 1996, Melting temperatures of the Earth’s mantle and core: Earth’s thermal structure: Annual Review of Earth and Planetary Sciences, v. 24, no. 1, p. 15–40., doi: 10.1146/annurev.earth.24.1.15.
  4. Bowen, N.L., 1922, The Reaction Principle in Petrogenesis: J. Geol., v. 30, no. 3, p. 177–198.
  5. Bowen, N.L., 1928, The evolution of the igneous rocks: Dover Publications, 334 p.
  6. Carr, M.H., 1975, Geologic map of the Tharsis Quadrangle of Mars: IMAP.
  7. Earle, S., 2015, Physical geology OER textbook: BC Campus OpenEd.
  8. EarthScope, 2014, Mount Ontake Volcanic Eruption: Online, http://www.earthscope.org/science/geo-events/mount-ontake-volcanic-eruption, accessed July 2016.
  9. Frankel, C., 2005, Worlds on Fire: Volcanoes on the Earth, the Moon, Mars, Venus and Io: Cambridge University Press, 396 p.
  10. Glazner, A.F., Bartley, J.M., Coleman, D.S., Gray, W., and Taylor, R.Z., 2004, Are plutons assembled over millions of years by amalgamation from small magma chambers? GSA Today, v. 14, no. 4, p. 4., DOI: 10.1130/1052-5173(2004)014<0004:APAOMO>2.0.CO;2
  11. Luongo, G., Perrotta, A., Scarpati, C., De Carolis, E., Patricelli, G., and Ciarallo, A., 2003, Impact of the AD 79 explosive eruption on Pompeii, II. Causes of death of the inhabitants inferred by stratigraphic analysis and areal distribution of the human casualties: J. Volcanol. Geotherm. Res., v. 126, no. 3–4, p. 169–200.
  12. Mueller, S., and Phillips, R.J., 1991, On the initiation of subduction: J. Geophys. Res. [Solid Earth], v. 96, no. B1, p. 651–665.
  13. Peacock, M.A., 1931, Classification of Igneous Rock Series: The Journal of Geology, v. 39, no. 1, p. 54–67.
  14. Perkins, S., 2011, 2010’s Volcano-Induced Air Travel Shutdown Was Justified: Online, http://www.sciencemag.org/news/2011/04/2010s-volcano-induced-air-travel-shutdown-was-justified, accessed July 2016.
  15. Peterson, D.W., and Tilling, R.I., 1980, Transition of basaltic lava from pahoehoe to aa, Kilauea Volcano, Hawaii: Field observations and key factors – ScienceDirect: J. Volcanol. Geotherm. Res., v. 7, no. 3–4, p. 271–293.
  16. Petrini and Podladchikov, 2000, Lithospheric pressure–depth relationship in compressive regions of thickened crust: Journal of Metamorphic Geology, v. 18, no. 1, p. 67–77., doi: 10.1046/j.1525-1314.2000.00240.x.
  17. Reid, J.B., Evans, O.C., and Fates, D.G., 1983, Magma mixing in granitic rocks of the central Sierra Nevada, California: Earth and Planetary Science Letters, v. 66, p. 243–261., doi: 10.1016/0012-821X(83)90139-5.
  18. Rhodes, J.M., and Lockwood, J.P., 1995, Mauna Loa Revealed: Structure, Composition, History, and Hazards: Washington DC American Geophysical Union Geophysical Monograph Series, v. 92.
  19. Scandone, R., Giacomelli, L., and Gasparini, P., 1993, Mount Vesuvius: 2000 years of volcanological observations: Journal of Volcanology and Geothermal Research, v. 58, p. 5–25.
  20. Stovall, W.K., Wilkins, A.M., Mandeville, C.W., and Driedger, C.L., 2016, Fact Sheet.
  21. Thorarinsson, S., 1969, The Lakagigar eruption of 1783: Bull. Volcanol., v. 33, no. 3, p. 910–929.
  22. Tilling, R.I., 2008, The critical role of volcano monitoring in risk reduction: Adv. Geosci., v. 14, p. 3–11.
  23. United States Geological Survey, 1999, Exploring the deep ocean floor: Online, http://pubs.usgs.gov/gip/dynamic/exploring.html, accessed July 2016.
  24. United States Geological Survey, 2012, Black Rock Desert Volcanic Field: Online, http://volcanoes.usgs.gov/volcanoes/black_rock_desert/, accessed July 2016.
  25. USGS, 2001, Dual volcanic tragedies in the Caribbean led to founding of HVO: Online, http://hvo.wr.usgs.gov/volcanowatch/archive/2001/01_05_03.html, accessed July 2016.
  26. USGS, 2011, Volcanoes: Principal Types of Volcanoes: Online, http://pubs.usgs.gov/gip/volc/types.html, accessed July 2016.
  27. USGS, 2012a, USGS: Volcano Hazards Program: Online, https://volcanoes.usgs.gov/vhp/hazards.html, accessed July 2016.
  28. USGS, 2012b, Yellowstone Volcano Observatory: Online, https://volcanoes.usgs.gov/volcanoes/yellowstone/yellowstone_geo_hist_52.html, accessed July 2016.
  29. USGS, 2016, Volcanoes General – What are the different types of volcanoes? Online, https://www2.usgs.gov/faq/categories/9819/2730, accessed March 2017.
  30. USGS, 2017, The Volcanoes of Lewis and Clark – Mount St. Helens: Online, https://volcanoes.usgs.gov/observatories/cvo/Historical/LewisClark/Info/summary_mount_st_helens.shtml, accessed March 2017.
  31. Wallace, P.J., 2005, Volatiles in subduction zone magmas: concentrations and fluxes based on melt inclusion and volcanic gas data: Journal of Volcanology and Geothermal Research, v. 140, no. 1–3, p. 217–240., doi: 10.1016/j.jvolgeores.2004.07.023.
  32. Williams, H., 1942, The Geology of Crater Lake National Park, Oregon: With a Reconnaissance of the Cascade Range Southward to Mount Shasta: Carnegie institution.

Figure References

Figure 4.1: Lava flow in Hawai’i. Brocken Inaglory. 2007. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:P%C4%81hoehoe_and_Aa_flows_at_Hawaii.jpg

Figure 4.2: Half Dome, an intrusive igneous batholith in Yosemite National Park. Jon Sullivan. 2004. Public domain. https://commons.wikimedia.org/wiki/File:Yosemite_20_bg_090404.jpg

Figure 4.3: Granite is a classic coarse-grained (phaneritic) intrusive igneous rock. James St. John. 2019. CC BY 2.0. https://commons.wikimedia.org/wiki/File:Granite_47_(49201189712).jpg

Figure 4.4: Basalt is a classic fine-grained extrusive igneous rock. James St. John. 2019. CC BY 2.0. https://commons.wikimedia.org/wiki/File:Basalt_3_(48674276863).jpg

Figure 4.5: Porphyritic texture. Jstuby. 2008. Public domain. https://commons.wikimedia.org/wiki/File:Olearyandesite.jpg

Figure 4.6: Pegmatitic texture. Jstuby. 2007. Public domain. https://commons.wikimedia.org/wiki/File:We-pegmatite.jpg

Figure 4.7: Scoria. Jonathan Zander (Digon3). 2008. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:Scoria_Macro_Digon3.jpg

Figure 4.8: Pumice. deltalimatrieste. 2008. Public domain. https://commons.wikimedia.org/wiki/File:Pomice_di_veglia.jpg

Figure 4.9: Obsidian (volcanic glass). Note conchoidal fracture. Ji-Elle. 2011. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:Lipari-Obsidienne_(5).jpg

Figure 4.10: Welded tuff. Wilson44691. 2010. Public domain. https://commons.wikimedia.org/wiki/File:HoleInTheWallTuff.JPG

Figure 4.11: Mineral composition of common igneous rocks. Woudloper. 2009. Public domain. https://commons.wikimedia.org/wiki/File:Mineralogy_igneous_rocks_EN.svg

Figure 4.12: Igneous rock classification table with composition as vertical columns and texture as horizontal rows. Kindred Grey. 2022. Adapted from Belinda Madsen, An Introduction to Geology. OpenStax. Salt Lake Community College. CC BY-NC-SA 4.0.

Table 4.1: Aphanitic and phaneritic rock types with images. Quartz monzonite 36mw1037 by B.W. Hallett, V. F. Paskevich, L.J. Poppe, S.G. Brand, and D.S. Blackwood via USGS (Public domain, https://commons.wikimedia.org/wiki/File:Quartz_monzonite_36mw1037.jpg). PinkRhyolite by Michael C. Rygel, 2014 (CC BY-SA 3.0, https://commons.wikimedia.org/wiki/File:PinkRhyolite.tif). Diorite MA by Amcyrus2012, 2015 (CC BY 4.0, https://commons.wikimedia.org/wiki/File:Diorite_MA.JPG). Andesite by James St. John, 2014 (CC BY 2.0, https://flic.kr/p/oBkKSy). GabbroRockCreek1 by Mark A. Wilson, 2008 (Public domain, https://commons.wikimedia.org/wiki/File:GabbroRockCreek1.jpg). VesicularBasalt1 by Jstuby, 2008 (Public domain, https://commons.wikimedia.org/wiki/File:VesicularBasalt1.jpg).

Figure 4.13: Dike of olivine gabbro cuts across Baffin Island in the Canadian Arctic. Mike Beauregard. 2012. CC BY 2.0. https://en.wikipedia.org/wiki/File:Franklin_dike_on_northwestern_Baffin_Island..jpg

Figure 4.14: Igneous sill intruding between Paleozoic strata in Nova Scotia. Mikenorton. 2010. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:Horton_Bluff_mid-Carboniferous_sill.JPG

Figure 4.15: Quartz monzonite in the Cretaceous of Montana, USA. James St. John. 2010. CC BY 2.0. https://commons.wikimedia.org/wiki/File:Butte_Quartz_Monzonite_(Late_Cretaceous,_76_Ma;_Rampart_Mountain,_northeast_of_Butte,_Montana,_USA)_1.jpg

Figure 4.16: Half Dome in Yosemite National Park, California, is a part of the Sierra Nevada batholith which is mostly made of granite. Jon Sullivan. 2004. Public domain. https://commons.wikimedia.org/wiki/File:Yosemite_20_bg_090404.jpg

Figure 4.17: The Henry Mountains in Utah are interpreted to be a laccolith, exposed by erosion of the overlying layers. Steven Mahoney. 2005. CC BY-SA 2.5. https://commons.wikimedia.org/wiki/File:Henry_Mountains,_Utah,_2005-06-01.jpg

Figure 4.18: Laccolith forms as a blister in between sedimentary strata. Erimus and Stannered. 2007. Public domain. https://commons.wikimedia.org/wiki/File:Laccolith.svg

Figure 4.19: Bowen’s Reaction Series. Colivine. 2011. Public domain. https://commons.wikimedia.org/wiki/File:Bowen%27s_Reaction_Series.png

Figure 4.20: Olivine, the first mineral to crystallize in a melt. S kitahashi. 2006. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:Peridot2.jpg

Figure 4.21: Norman L. Bowen. Unknown author. 1909. Public domain. https://commons.wikimedia.org/wiki/File:NormanLBowen_1909.jpg

Figure 4.22: Norman L. Bowen and his colleague working at the Carnegie Institution of Washington Geophysical Laboratory. Smithsonian Institution. 2010. Public domain. https://commons.wikimedia.org/wiki/File:(left_to_right)-_Norman_Levi_Bowen_(1887-1956)_and_Orville_Frank_Tuttle_(1916-1983)_(4730112454)_(cropped).jpg

Figure 4.23: Geothermal gradient. Bkilli1. 2013. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:Temperature_schematic_of_inner_Earth.jpg

Figure 4.24: Pressure-temperature diagram showing temperature in degrees Celsius on the x-axis and depth below the surface in kilometers (km) on the y-axis. Kindred Grey. 2022. CC BY-SA 3.0. Adapted from Partial melting asthenosphere EN by Woudloper, 2010 (CC BY-SA 3.0, https://commons.wikimedia.org/wiki/File%3APartial_melting_asthenosphere_EN.svg).

Figure 4.25: Association of volcanoes with plate boundaries. Jose F. Vigil via USGS. 1997. Public domain. https://commons.wikimedia.org/wiki/File:Tectonic_plate_boundaries.png

Figure 4.26: Map of spreading ridges throughout the world. Eric Gaba. 2006. CC BY-SA 2.5. https://commons.wikimedia.org/wiki/File:Spreading_ridges_volcanoes_map-fr.svg

Figure 4.27: Pillow basalt on sea floor near Hawai’i. NOAA. 1988. Public domain. https://commons.wikimedia.org/wiki/File:Nur05018-Pillow_lavas_off_Hawaii.jpg

Figure 4.28: Black smoker hydrothermal vent with a colony of giant (6’+) tube worms. NOAA. 2006. Public domain. https://commons.wikimedia.org/wiki/File:Main_Endeavour_black_smoker.jpg

Figure 4.29: Distribution of hydrothermal vent fields. DeDuijn. 2016. CC BY-SA 4.0. https://commons.wikimedia.org/wiki/File:Distribution_of_hydrothermal_vent_fields.png

Figure 4.30: Distribution of volcanoes on the planet. USGS. 2007. Public domain. https://commons.wikimedia.org/wiki/File:Map_plate_tectonics_world.gif

Figure 4.31: Basaltic cinder cones of the Black Rock Desert near Beaver, Utah. Lee Siebert via Smithsonian Institution. 1996. Public domain. https://commons.wikimedia.org/wiki/File:Black_Rock_Desert_volcanic_field.jpg

Figure 4.32: Diagram showing a non-moving source of magma (mantle plume) and a moving overriding plate. Los688. 2008. Public domain. https://en.wikipedia.org/wiki/File:Hotspot(geology)-1.svg

Figure 4.33: The track of the Yellowstone hotspot, which shows the age of different eruptions in millions of years ago. Kelvin Case. 2013. CC BY 3.0. https://commons.wikimedia.org/wiki/File:HotspotsSRP_update2013.JPG

Figure 4.34: The Hawaiian–Emperor seamount and island chain. Ingo Wölbern. 2008. Public domain. https://commons.wikimedia.org/wiki/File:Hawaii-Emperor_engl.png

Figure 4.35: Mt. Shasta in Washington state with Shastina, its parasitic cone. Don Graham. 2013. CC BY-SA 2.0. https://commons.wikimedia.org/wiki/File:Mt._Shasta_and_Mt._Shastina,_CA_9-13_(26491330883).jpg

Figure 4.36: Oregon’s Crater Lake was formed about 7700 years ago after the eruption of Mount Mazama. Zainubrazvi. 2006. CC BY-SA 3.0. https://en.wikipedia.org/wiki/File:Crater_lake_oregon.jpg

Figure 4.37: Kilauea in Hawai’i. Quinn Dombrowski. 2007. CC BY-SA 2.0. https://commons.wikimedia.org/wiki/File:Kilauea_Shield_Volcano_Hawaii_20071209A.jpg

Figure 4.38: Eruption of Kilauea in 2018 produced high viscosity lava shown here crossing a road. USGS. 2018. Public domain. https://commons.wikimedia.org/wiki/File:USGS_K%C4%ABlauea_multimediaFile-1955.jpg

Figure 4.39: Olympus Mons, an enormous shield volcano on Mars, the largest volcano in the solar system, standing about two and a half times higher than Everest is above sea level. NASA. 1978. Public domain. https://en.wikipedia.org/wiki/File:Olympus_Mons_alt.jpg

Figure 4.40: Ropey pahoehoe lava. Bbb. 2010. GNU Free Documentation License 1.2. https://de.wikivoyage.org/wiki/Datei:ReU_PtFournaise_Lavastr%C3%B6me.jpg

Figure 4.41: Blocky a’a lava. Librex. 2009. CC BY 2.0. https://commons.wikimedia.org/wiki/File:Lava_del_Volcan_Pacaya_2009-11-28.jpg

Figure 4.42: Volcanic fissure and flow, which could eventually form a lava tube. NPS. 2004. Public domain. https://commons.wikimedia.org/wiki/File:Volcano_q.jpg

Figure 4.43: Devils Tower in Wyoming has columnar jointing. Colin.faulkingham. 2005. Public domain. https://commons.wikimedia.org/wiki/File:Devils_Tower_CROP.jpg

Figure 4.44: Columnar jointing on Giant’s Causeway in Ireland. Udri. 2014. CC BY-NC-SA 2.0. https://flic.kr/p/2j1mgwE

Figure 4.45: Mount Rainier towers over Tacoma, Washington. Lyn Topinka via USGS. 1984. Public domain. https://commons.wikimedia.org/wiki/File:Mount_Rainier_over_Tacoma.jpg

Figure 4.46: Mt. Fuji in Japan, a typical stratovolcano, symmetrical, increasing slope, visible crater at the top. Alpsdake. 2016. CC BY-SA 4.0. https://commons.wikimedia.org/wiki/File:Numazu_and_Mount_Fuji.jpg

Figure 4.47: Lava domes have started the rebuilding process at Mount St. Helens, Washington. Willie Scott via USGS. 2006. Public domain. https://commons.wikimedia.org/wiki/File:MSH06_aerial_crater_from_north_high_angle_09-12-06.jpg

Figure 4.48: Timeline of events at Mount Mazama. USGS and NPS. 2006. Public domain. https://commons.wikimedia.org/wiki/File:Mount_Mazama_eruption_timeline.PNG

Figure 4.49: Wizard Island sits in the caldera at Crater Lake. Don Graham. 2006. CC BY-SA 2.0. https://commons.wikimedia.org/wiki/File:Crater_Lake_National_Park,_OR_2006_(6539577313).jpg

Figure 4.50: Map of calderas and related rocks around Yellowstone. USGS. 1905. Public domain. https://www.usgs.gov/media/images/simplified-map-yellowstone-caldera

Figure 4.51: Several prominent ash beds found in North America, including three Yellowstone eruptions shaded pink (Mesa Falls, Huckleberry Ridge, and Lava Creek), the Bisho Tuff ash bed (brown dashed line), and the modern May 18th, 1980 ash fall (yellow). USGS. 2005. Public domain. https://commons.wikimedia.org/wiki/File:Yellowstone_volcano_-_ash_beds.svg

Figure 4.52: Sunset Crater, Arizona is a cinder cone. NPS. Unknown date. Public domain. https://commons.wikimedia.org/wiki/File:Sunset_Crater10.jpg

Figure 4.53: Soon after the birth of Parícutin in 1943. K. Segerstrom via USGS. 1943. Public domain. https://commons.wikimedia.org/wiki/File:Paricutin_30_612.jpg

Figure 4.54: Lava from Parícutin covered the local church and destroyed the town of San Juan, Mexico. Sparksmex. 2007. Public domain. https://commons.wikimedia.org/wiki/File:Paricutin2.jpg

Figure 4.55: World map of flood basalts. Williamborg. 2011. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:Flood_Basalt_Map.jpg

Figure 4.56: Igneous rock types and related volcano types. Unknown author. Unknown date. “Personal use.” https://www.seekpng.com/ipng/u2t4y3r5o0r5r5w7_table-of-igneous-rocks-and-related-volcano-types/

Figure 4.57: General diagram of volcanic hazards. USGS. 2008. Public domain. https://pubs.usgs.gov/fs/fs002-97/old%20html%20files/

Figure 4.58: Human remains from the 79 CE eruption of Vesuvius. Gary Todd. 2019. Public domain. https://commons.wikimedia.org/wiki/File:Pompeii_Ruins_Cast_of_Human_Victim_at_Villa_of_the_Mysteries_(48445486616).jpg

Figure 4.59: Mount St. Helens, the day before the May 18th, 1980 eruption (left) and 4 months after the major eruption (right). Mount St. Helens, one day before the devastating eruption by Harry Glicken, USGS/CVO, 1980 (Public domain, https://commons.wikimedia.org/wiki/File:Mount_St._Helens,_one_day_before_the_devastating_eruption.jpg). MSH80 st helens from johnston ridge 09-10-80 by Harry Glicken via USGS (Public domain, https://commons.wikimedia.org/wiki/File:MSH80_st_helens_from_johnston_ridge_09-10-80.jpg).

Figure 4.60: Image from the May 18, 1980, eruption of Mt. Saint Helens, Washington. Austin Post via USGS. 1980. Public domain. https://commons.wikimedia.org/wiki/File:MSH80_eruption_mount_st_helens_05-18-80-dramatic-edit.jpg

Figure 4.61: The material coming down from the eruption column is a pyroclastic flow. C.G. Newhall via USGS. 1984. Public domain. https://en.wikipedia.org/wiki/File:Pyroclastic_flows_at_Mayon_Volcano.jpg#:~:text=English%3A%20Pyroclastic%20flows%20at%20Mayon,50%20km%20toward%20the%20west.

Figure 4.62: The remains of St. Pierre. Angelo Heilprin. 1902. Public domain. https://commons.wikimedia.org/wiki/File:Pelee_1902_3.jpg

Figure 4.63: Sequence of events for Mount St. Helens, May 18th, 1980. Lyn Topinka via USGS. 1998. Public domain. https://commons.wikimedia.org/wiki/File:Msh_may18_sequence.gif

Figure 4.64: Aman sweeps ash from an eruption of Kelud, Indonesia. Crisco 1492. 2014. CC BY-SA 3.0. https://commons.wikimedia.org/wiki/File:Ash_in_Yogyakarta_during_the_2014_eruption_of_Kelud_01.jpg

Figure 4.65: Micrograph of silica particle in volcanic ash. USGS. 1980. Public domain. https://volcanoes.usgs.gov/volcanic_ash/components_ash.html

Figure 4.66: Mud line shows the extent of lahars around Mount St. Helens. USGS. 1980. Public domain. https://www.usgs.gov/media/images/lahars-resulting-may-18-1980-eruption-mount-st-helens

Figure 4.67: Old lahars around Tacoma, Washington. USGS. 1905. Public domain. https://www.usgs.gov/media/images/lahar-pathways-events-heading-mount-rainier-map-showing-t

 

definition

License

Icon for the Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License

Introduction to Earth Science Copyright © 2023 by Laura Neser is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License, except where otherwise noted.

Share This Book