Analcime |Na(H2O)| [AlSi2O6]
     
Morphology:  
  Isometric or pseudo-isometric
Single crystals are trapezohedra in sizes ranging from millimeters to several centimeters
Common forms: {211} and rarely {100}
 
Physical properties:
  Cleavage:  {100} poor
Hardness: 5 –5.5
Density:  2.22 g/cm3
Luster:  vitreous
Streak:  white
  Analcime (largest crystal 2 cm) with serandite and aegerine, Mont Saint-Hilaire, Quebec, Canada
Optical properties:  
  Color:  colorless to gray
colorless in thin section
Isotropic or very weakly anisotropic
n = 1.479 – 1.493
      Analcime (largest crystal 60 μm) in the Barstow Formation, San Bernardino County, California, USA. (Sheppard and Gude 1969)
Crystallography:  
  Unit cells
  isometric a = 13.725 Å, space group Ia3d
  tetragonal a = 13.723 Å ,   c = 13.686 Å, space group I41/acd or I41/a
a = 13.721 Å,    c = 13.735 Å
  orthorhombic a = 13.733 Å,  b = 13.729 Å   c = 13.712 Å, space group Ibca
a
= 13.727 Å,  b = 13.714 Å   c = 13.740 Å
  monoclinic a = 13.73 Å,  b = 13.72 Å   c = 13.74 Å, β = 90.0 - 90.5°, space group I2/a
  triclinic a = 13.6824 Å,  b = 13.7044 Å   c = 13.7063 Å,
α = 90.158°, β = 89.569°, γ = 89.545°, space group P1
(Mazzi and Galli, 1978)
  Z = 16  
     
Name:  
  Haüy (1797, 1801) gave the name analcime to the mineral that had been called by various informal names, such as zeolithe dure. The new name is derived from the Greek word for “without strength” in allusion to its weak electrical effects induced by friction. Haüy (1801) assigned the occurrence near Catania on the Isola dei Ciclopi (Cyclopean Islands), Sicily, Italy, as the type locality. Analcime is a common zeolite occurring in cavities of altered mafic volcanic rocks, possibly as a primary phase in some undersaturated volcanic rocks, a common product of burial metamorphism, a diagenetic mineral replacing various materials in both open and closed hydrologic systems, precipitates from strongly alkaline waters, and in low temperature hydrothermal veins.
       
Crystal structure:  
  The determination of the structure of analcime by Taylor (1930) was a milestone in zeolite mineralogy. Not only was it the first zeolite structure to be solved, but the discovery that (Si,Al)O4 tetrahedra are in a framework arrangement became the basis for a new definition of zeolite (Hey 1930). Taylor (1930) determined the structure in the cubic space group Ia3d.
 

The analcime framework consists of singly-connected 4-rings, arranged in chains coiled around tetrad screw axes (see ANA and Gottardi and Galli, 1985, p. 9).  Parallel chains alternate 41 and 43 screw axes. Every 4-ring is a part of three mutually perpendicular chains, each parallel to a crystallographic axis. Cages, which contain the Na-cations and water molecules, occur near where chains interconnect, and each T-site is part of three cages. In the cubic space group every T-site is equivalent to every other T-site, and therefore, the Si,Al distribution among these sites must be random. Na

  cations (yellow in the figure) are in the centers of these cages, but there are 24 cages in the unit cell. Therefore, the Na cations (generally about 16, but may be from 12 to 17) must also be randomly distributed among  the cages. Water molecules (blue circles) fully occupy the 16 sites in the unit cell. Any excess water molecules must be randomly distributed in unoccupied Na sites.
  It has long been known that many analcime crystals are not optically isotropic, and therefore, must be non-cubic. Early single crystal X-ray diffraction work by Coombs (1955) showed that such crystals are indeed non-cubic. Mazzi and Galli (1978) refined the structures of seven different analcime crystals, five in the tetragonal space group I41/acd and two in the orthorhombic space group Ibca. Most tetragonal crystals have a > c, but one sample has c > a. Hazen and Finger (1979) determined the unit cell dimensions of another different 15 analcime crystals, many of which are cubic, tetragonal, and orthorhombic. Several have monoclinic symmetry with the 2-fold axes parallel to either the pseudocubic [100] or [110] directions, and one is triclinic.
  Lower symmetry is a result of preferential Al occupancy in some of the T-sites. Mazzi and Galli (1978) showed that in the tetragonal crystals with a > c two cages are equivalent and contain most of the Na with Al concentrated in the nearby T-sites. For tetragonal crystals with c > a the two equivalent cages have less Na and associated Al than the third. In orthorhombic crystals all three cages have different Na occupancies and Al in the nearby T-sites. Because the structure of monoclinic and triclinic crystals have not been refined, ordering patterns in these crystal are not known, but the ordering may be similar to that in wairakite. Fine lamellar, pseudo-merohedral twinning on {110} is present in all non-cubic crystals.
   
Chemical composition:
  The compositional range of analcime can be expressed with the generalized formula: 
|Na16-x (H2O)16+x| [Al16-x Si32+x O96], in which x varies between -3.4 and +4.3. Representative analyses are given in Deer et al. (2004, table 24). The main compositional variation is in the Si/Al of the framework with the necessary adjustments of Na for charge balance and possibly H2O to make room for Na or to fill empty sites. Over all analyzed samples Si ranges from 28.6 (approximately equivalent to the Si/Al of natrolite) to 36.3 (similar to albite) per unit cell. To a large extent the Si content is related to the precursor materials (glass or zeolite) and the environment of crystallization. With the exception of samples from some metamorphic rocks, analcime has very minor amounts of non-framework cations other than Na.
  All early analyses and many of the more recent ones are on coarse-grained crystals from cavities in basaltic rocks. Compositions in this group tend to cluster near 32 Si per unit cell, but vary between 31 and 33. Analcime from all other occurrences shows a wider variation. The lowest Si contents are from analcime that crystallized in igneous rocks, where there is some evidence suggesting that analcime might be a liquidus phase or in other cases is a result of subsolidus reactions. Most of these samples contain between 0.6 to 0.9 wt % Fe2O3, which may be from inclusions, but the reasonable charge balances in some cases suggest the Fe3+ is actually in tetrahedral sites.
  Because analcime has the same framework as wairakite, leucite, pollucite, and ammonioleucite, one might expect the major cations of these minerals, Ca, K, Cs, and NH4, respectively, to be present in at least moderate amounts. However, this is true only in rare circumstances.
  Very low-grade metamorphic rocks, whether formed by burial or weak contact metamorphism, have the highest Ca contents. Surdam (1966) and Harada et al. (1972) report samples intermediate between analcime and wairakite. In all metamorphic rocks Si varies between 31.7 and 34.5 atoms per unit cell, including samples surveyed by the X‑ray method (Coombs and Whetten, 1967).  Analcime that formed by the replacement of leucite commonly have high K (greater than about 0.7 ions per unit cell), that may be residual from precursor crystal.
  Analcime as an alteration product of terrestrial accumulations of rhyolitic tuff and ignimbrite in both open and closed hydrologic systems tends to have compositions between 34 and 35 Si per unit cell. The highest Si contents known are from active hydrothermal systems of Yellowstone National Park, where contents over 36 Si per unit cell have been measured. The Yellowstone samples contain very little Ca, but have fractionated Cs (up to 4700 ppm) from the fluids.
  The water content of analcime, especially those with high Si compositions, is not clearly defined by available analytical data. Even with the more recent analyses given in Deer et al. (2004, table 24), there is a wide scatter of points, indicating a substantial variation in composition in samples with Si contents near 32 per unit cell. There is also a general increase of water content along the trend toward 35 Si per unit cell. Sixteen water molecules per unit cell corresponds to 8.18 weight per cent H2O, and 17 corresponds to 8.65 weight per cent H2O. Therefore, samples must be carefully cleaned of included minerals (clays), equilibrated with a constant humidity, and analyzed using as large a sample as possible. Estimating water content by difference, as with microprobe analyses, will not yield useful results.
       
Occurrences:
  Analcime is a widespread zeolite occurring in many different kinds of environments, including amygdaloidal filling of cavities in basaltic rocks, replacement of rhyolitic vitric tuff, and precipitates from strongly alkaline waters. It occurs in a few igneous rocks, possibly as a primary mineral, and is an important constituent of volcaniclastic sediment that has been affected by burial metamorphism or by hydrothermal alteration. The following is largely adapted from Deer et al. (2004, p. 546-560).
  Diagenesis and burial metamorphism of sediment and sedimentary rocks
    In terrestrial accumulations of volcaniclastic sediment and rock, analcime is a common alteration product in various kinds of sediment and soil in hydrologically closed systems and in thick sections of tephra and ignimbrite in hydrologically open systems. Analcime replaces early-formed clinoptilolite in marine, andesitic volcaniclastic sediment of fore- and back-arc basins, persisting into burial metamorphic conditions.
  Hydrologically closed systems - tuff in lacustrine sediment
    The water of lakes in closed basins in arid climates tends to become alkaline with high salinity. Zeolites are among the authigenic minerals to crystallize in this environment, especially as a replacement of vitric tuff incorporated into the lacustrine sediment. Analcime in these lake sediments commonly replaces early-formed zeolites, such as clinoptilolite and phillipsite, but may also directly replace glass. There is evidence that analcime may precipitate directly from saline and alkaline water associated with non-tuffaceous sediment.
    One of the first descriptions of analcime forming major portions of altered tuff beds is from the Big Sandy Formation near Wikieup, Arizona (Ross, 1928; Sheppard and Gude, 1973).
    Other examples are the basins of the Anatolia region of Turkey (Echle, 1975; Ataman and Beseme, 1972; Gündogdu et al., 1996; Whateley et al., 1996). Some basins in the Ankara area have beds of analcime-carbonate argillite (Ataman and Gündogdu, 1982), which may occur as chemical precipitates, similar to the formation of analcimolite beds, see below.
    Tertiary lacustrine deposits are widespread in Serbia (Obradovic et al., 1995; Obradovic, 1988; Gottardi and Obradovic, 1978) and range from rhyodacite to andesite. In the Slanci basin near Belgrade authigenic minerals are horizontally zoned from fresh glass to montmorillonite to clinoptilolite to analcime (Obradovic, 1988). Similar deposits in Russia have been reported in supersaline deposits of the Solikamsk Basin (Polikarpov et al., 1986).
    Several basins in Argentina consist of lacustrine sediment that includes zeolitized tuff and tuffaceous sandstone beds. Of these, analcime occurs in claystone in the Tertiary Rio Sali Formation in the Tucuman Province (Teruggi and Andreis, 1963). The San Jorge Gulf Basin is filled with continental sediment, and the Cretaceous part of the fill comprises the Chubutian Group. Teruggi (1964) describes the Castillo Formation from this group, and the analcime replacement of tuff and tuffaceous sandstone. Analcime is the major authigenic mineral in the tuff of the Triassic Dark Victor Formation at Tupungato, Mendoza (Baldwin, 1944).
    Analcime occurs as a diagenetic mineral in oil producing sands of the Junggar Basin, an             important oil-producing region in northwest China (Tang et al., 1997).
  Hydrologically closed systems - lacustrine sediment with no volcanic glass
    In closed basins the development of high salinity and alkalinity along the margins of lakes or in mudflats provides the chemical environment for analcime formation without the involvement of volcanic material. The resulting rocks include analcimolite, extensive beds of nearly pure analcime, and ferruginous, analcimic argillite or shale. On the land surface, salinity and alkalinity in soil is also conducive to analcime development. Analcime forming reactions appear to involve Na-rich water, high pH, and an alumina and silica source, such as kaolinite or smectite. The composition of analcime formed in this manner tends to have a lower Si content (about 33.5 Si per unit cell) than that formed as a replacement of rhyolitic glass (34.5 to 35.0 Si per unit cell). In Lake Natron, Tanzania, present day formation of analcime occurs in non-tuffaceous clay in the lake bottom sediment (Hay, 1966, p. 36-39).
    Other examples include the three lithofacies associated with a Pleistocene lake in Olduvai Gorge, Tanzania  (Hay, 1966, 1970) the present day Lake Natron (Hay, 1970), the diagenetic alteration of lake-margin sediment on the northern margin of the saline, alkaline Lake Bogoria in the Kenya Rift Valley is a recent analogue of the development of reddish-brown, analcime-bearing argillite (Renaut, 1993), lake-margin sediment that formed in the Unita Basin of the Green River Formation, Utah, USA (Remy and Ferrell, 1989), the analcime-containing lake and lake-margin sediments of the Carboniferous Rocky Brook Formation, western Newfoundland, Canada (Gall and Hyde, 1989), and the abundant authigenic analcime in the Triassic Lockatong Formation deposited in the Newark Basin, New Jersey (Van Houten, 1960, 1962, 1965), and similar sequences in the Hartford  and Deerfield Basins in Connecticut and Massachusetts (van de Kamp and Leake, 1996).
    Analcimolite, thick beds (several tens of meters) of nearly pure analcime with wide areal extent (hundreds of km), is known from several localities and all appear to have formed from lake-margin mudstone associated with saline lakes. The Triassic Pogo Agie Member of the Chugwater Formation in Wyoming, USA, consists largely of ferruginous analcimolite with siltstone and claystone (Keller, 1952; High and Picard, 1965). Ferruginous analcimolite occurs in the Triassic Chinle Formation of Utah, USA, (Keller, 1953). About 40 m of Lower Cretaceous analcimolite occurs in the Aïr region, Niger (Joulia et al., 1959). The Upper Jurassic and Lower Cretaceous analcimolite beds of the Congo Basin in the Republic of the Congo are tens of meters thick and part of an 800 m section of analcimic argillite (Vanderstappen and Verbeek, 1959, 1964).
  Soil and surficial deposits
    Analcime occurs in soils of arid environments in settings like those of the margins of saline and alkaline lakes. Reported occurrences are in California, USA, east Africa, Russia, and India. Such soils are in areas that are poorly drained and develop local accumulations of sodium carbonate. Analcime and illite are found in all soil samples with excess alkali, i.e. accumulated sodium carbonate. The source of silica and alumina is most likely the abundant montmorillonite.
  Hydrologically open systems
    Terrestrial accumulations of pyroclastic debris, especially rhyolitic tephra and ignimbrite units, may alter to produce zeolites, particularly analcime. Because the zeolites occur largely from reactions with through flowing vadose and groundwater, this type of process is called hydrologically open alteration (Hay and Sheppard, 1977; Sheppard and Hay, 2001). Analcime occurs in intermediate zones, where it either replaces earlier formed zeolites or replaces glass shards directly. Alteration zones are oriented vertically and are controlled largely by temperature. A well-investigated example of the alteration of a thick pile of pyroclastic rocks is Yucca Mountain along the southwest edge of the Nevada Test Site in Nye County, Nevada, USA (Broxton et al., 1987).
  Deep sea sediment
 

 

Analcime is a component of the authigenic minerals of deep-sea sediment, but is much less common than clinoptilolite or phillipsite. It was first reported in deep-sea sediment by Murray and Renard (1891), who noted that it is always associated with basaltic glass. Iijima (1978) and Kastner and Stonecipher (1978), reviewing the occurrence of analcime in cores from the Deep Sea Drilling Project, note it tends to be in volcanic sediment. There is no evidence that analcime replaces precursor zeolite, such as clinoptilolite. Iijima (1978) comments that evidence available at the time of his review did not indicate whether analcime precipitated from cold interstitial water or from hydrothermal alteration. Further drilling has revealed an environment, trench margins, where analcime is produced from heating by intrusion of dikes and sills into the sedimentary column.
  Diagenesis and burial metamorphism of marine sediment from arc-source terrains
    Analcime is a constituent, although not a major one, of authigenic minerals in volcaniclastic sediment in basins near volcanic arcs. With rapid deposition and burial of andesitic volcaniclastic sediment, diagenesis merges into burial metamorphism. The importance of zeolites in these rocks was first recognized in a thick section exposed in the Taringatura Hills, New Zealand by Coombs (1954). The concept of the Zeolite Facies of metamorphism, as defined by Turner (in Fyfe et al., 1958), was applied to several thick sequences of volcaniclastic rocks in Japan (Seki, 1969), Puerto Rico (Otálora, 1964), and the Hokonui Hills, New Zealand (Boles and Coombs, 1975). In these areas heulandite-Ca, clinoptilolite-Na, or clinoptilolite-Ca initially replace the glass of tuffaceous rocks. At deeper levels these minerals are generally replaced by analcime or authigenic feldspars. In some areas, such as Japan, vertical zonation is distinct, but in others breakdown of early phases occurs over wide stratigraphic intervals.
  Cavities in basaltic lava flows
 

 

Analcime is common in basalt cavities and is most prevalent in saturated to undersaturated basaltic host rocks. Some of the notable localities are the type locality on Cyclopean Islands (Isola dei Ciclopi), Sicily, Italy; at Mazé in the Iwamure district, Niigata Prefecture, Japan; Berufjord, Iceland; Val di Fassa, Alpe di Siusi, Italy; Dumbarton and Salisbury Crags, Scotland; Table Mountain, Colorado; Skookumchuck Dam, Tenino, Washington; and Paterson, New Jersey, USA. Common associated minerals include natrolite, heulandite, chabazite, stilbite, laumontite, apophyllite, calcite, and quartz.
    In the active geothermal areas of Iceland analcime occurs over a wide range of temperatures (70° to 300°C). This suggests that analcime precipitation from solutions is controlled by variables other than temperature, such as silica activity, playing major roles in nucleation and stability.
  Analcime as a product of hydrothermal alteration
    Analcime crystallizes in hydrothermal systems, both in environments with moderately high water to rock ratios, geothermal systems, and in some with low ratios, such as end stage crystallization or deuteric alteration in syenite intrusions.
  Active geothermal systems
    In cores from the 15 research holes in the rhyolitic host rocks of the thermal areas of Yellowstone National Park, Wyoming, analcime occurs in C-1, Y-1, Y-2, Y-3, and Y-8. In the Y-3 hole analcime was found in drill-core from the temperature interval 90° to 140°C and from depths between 15 and 30 m (Bargar and Beeson, 1985). Published compositions show a range in Si content between 33.84 and 36.33 atoms per unit cell (Bargar and Beeson, 1985). Trace element analyses of analcime from several drill holes indicate Cs concentrations up to 4700 ppm (Keith et al., 1983).
    Koshiya et al. (1994) describe the zonal distribution of hydrothermal vein fillings in the Takinoue geothermal area, northern Honshu, Japan. Analcime occurs in the outermost zone, presumably forming at lower temperatures than the laumontite.
  Fossil hydrothermal areas
    Analcime is very common in the hydrothermal aureoles surrounding Kuroko-type deposits in northern Honshu, Japan (Utada, 1988), and is the index mineral of the analcime zone. The outermost is the mordenite zone, and analcime is next zone inward.
  Late stage, deuteric alteration of plutonic rocks
    Analcime crystallizes in miarolitic cavities and fractures in nepheline syenite, nepheline phonolite, and syenite pegmatite dikes and lenses. It may crystallize from late alkalic fluids or by reaction of fluids with nepheline. An example of this type of paragenesis is at Mont Saint-Hilaire, Quebec, where analcime occurs in unaltered pegmatite (initial crystallization), altered pegmatite (late fluid alteration), miarolitic cavities in nepheline syenite, and along breccia zones (Horvath and Gault, 1990). Crystals as large as 25 cm have crystallized in some breccia cavities, where it is associated with serandite and natrolite.
    In the Lovozero alkaline massif, Russia, analcime is wide-spread and common in pegmatite and hydrothermal bodies, especially those derived from the poikilitic syenite and from the differentiated intrusive complex (Pekov, 2000). The largest crystals, up to 13 cm, are in the Nastrophitovoye pegmatite on Alluaiv Mountain. Tranparent, pink crystals from naujaite of Alluaiv Mtn. are cubic; a yellow variety from foyaite pegmatite of the Kedykverpakhk Mtn, is tetragonal; and the crystals from the Nastrophitovoye pegmatite are orthorhombic (Pekov, 2000).
    Analcime is common in other syenite pegmatite dikes where it is associated with natrolite. The dikes in the Oslo region of Norway are good examples (Raade et al., 1983). Most miarolitic cavities of the phonolite sill at Point of Rocks, Colfax County, New Mexico, USA, contain analcime with albite, nepheline, and aegerine (DeMark, 1984). Henderson and Gibb (1983) investigated the occurrence of analcime in several differentiated mafic sills, including olivine theralite, Otago, New Zealand; crinanite, Madiera; crinanite, Howford Bridge sill, Ayrshire, Scotland; theralite, Lugar sill, Ayrshire, Scotland; essexite, Lennoxtown, Scotland; essexite and theralite, Papeno Valley, Tahiti; and nepheline “dolerite”, Loebauerberg and Heidelberg, Germany. They conclude that clear, interstitial analcime forms by low-temperature alteration of primary SiO2-rich nepheline.
  Analcime as a primary igneous mineral
 

 

Analcime occurs as phenocrysts and in the groundmass of some alkaline lavas and intrusive rocks, such as phonolite and nepheline syenite. Petrographic evidence mostly from volcanic rocks led many workers to the conclusion that analcime crystallized from magma of the appropriate composition (Harker, 1954; Wilkinson, 1965, 1968; Pearce, 1970, 1993; Goble et al., 1993; among others). These interpretations require analcime to be stable from late magmatic stages to deuteric conditions, but more recent experimental work shows analcime can be a liquidus phase only within a narrow temperature range at high fluid pressures (Roux and Hamilton, 1976). Furthermore, analcime is easily formed by reaction of certain fluids with nepheline or leucite (Saha, 1959; Gupta and Fyfe, 1975). Therefore, many recent studies of analcime-phyric volcanic rocks have come to the conclusion that the analcime phenocrysts are replacements of primary leucite, although without universal agreement.
    In Italy analcime replacement of leucite is widespread in the potassic lavas in the region around Lago di Vico (northwest of Rome). Cundari and Graziani (1964) have documented the process in tephritic lavas, some of which contain partially replaced leucite phenocrysts. Luhr and Giannetti (1987) describe analcime that has completely replaced centimeter-sized leucite phenocrysts in 385 ka leucite tephrite pyroclastic flow deposits from the Quaternary volcano, Roccamonfina (northwest of Naples).
       
Uses:  
  Because of its compact structure and the consequent strong resistance to diffusion of either molecules or cations, applications of analcime as microporous material are lacking. Nevertheless analcime-bearing tuffs are known to be sporadically employed as building materials. Use of analcime-rich tuffs as dimension stone or as pozzolanic addition to cement and concrete is, for instance, a consolidated practice in Bulgaria. Similar, but more accidental, uses are also reported in Georgia, Germany and Italy (in the latter case limited to dimension stone) (Colella et al., 2001; Colella, 2007).
       
References:  
  Ataman, G. and Beseme, P. (1972) Découverte de l’analcime sédimentaire en Anatolie du nord-ouest (Turquie): minéralogie, genèse, paragènese. Chem. Geol. 9, 203-225.

Ataman, G. and Gündogdu, N. (1982) Analcimic zones in the Tertiary of Anatolia and their geologic positions. Sediment. Geol. 31, 89-99.

Baldwin, J.L. (1944) Tupungate oil field, Mendoza, Argentina. Amer. Assoc. Petrol. Geol. Bull. 28, 1455-1484.

Bargar, K.E. and Beeson, M.H. (1985) Hydrothermal alteration in Research Drill Hole Y-3, Lower Geyser Basin, Yellowstone National Park, Wyoming. U.S. Geol. Surv., Prof. Paper  1054-C, 23 pp.

Boles, J.R. and Coombs, D.S. (1975) Mineral reactions in zeolitic Triassic tuff, Hokonui Hills, New Zealand. Geol. Soc. Amer., Bull. 86, 163-173.

Broxton, D.E., Bish, D.L., and Warren, R.G. (1987) Distribution and chemistry of diagenetic minerals at Yucca Mountain, Nye County, Nevada. Clays and Clay Miner. 35, 89-110.

Colella, C., de’ Gennaro, M,. and Aiello, R. (2001) Use of zeolitic tuff in the building industry. In: Bish, D.L. and Ming, D.W. (eds.) Natural Zeolites: Mineralogy, Occurrence, Properties, Applications, Reviews in Mineralogy & Geochemistry, Mineralogical Society of America, Vol. 45, Washington, D.C., 551-587.

Colella, C. (2007) Natural zeolites and environment. In: Čejka, J., van Bekkum, H., Corma, A. and Schueth, F. (eds.) Introduction to Zeolite Science and Practice. 3rd Revised Edition, Studies in Surface Science and Catalysis No. 168, Elsevier, Amsterdam, 999-1035.

Coombs, D.S. (1954) The nature and alteration of some Triassic sediments from Southland, New Zealand. Trans. Roy. Soc. N.Z. 82, 65-109.

Coombs, D.S. (1955) X-ray observations of wairakite and non-cubic analcime. Mineral. Mag. 30, 699-708.

Coombs, D.S. and Whetten, J.T. (1967) Composition of analcime from sedimentary and metamorphic rocks. Geol. Soc. Amer., Bull. 78, 269-282.

Cundari, A. and Graziani, G. (1964) Prodotti di alterazione della leucite nelle vulcaniti vicane. Period. Miner. 33, 35-52.

Deer, A., Howie, R., Wise, W.S., and Zussman, J. (2004) Rock Forming Minerals. vol. 4B. Framework Silicates: Silica Minerals, Feldspathoids and the Zeolites. The Geological Society, London.

DeMark, R.A. (1984) Minerals of Point of Rocks, New Mexico. Mineral. Rec. 15, 149-156.

Echle, W. (1975) Zusammensetung und Entstehung sedimentärer Analcime in Jungtertiären pyroklastischen Gestein nördlich Mihaliççik, Westanatolien, Türkei. Neues Jahrb. Miner. Abh. 124, 128-146.

Fyfe, W.S., Turner, F.J., and Verhoogen, J. (1958) Metamorphic reactions and metamorphic facies. Geol. Soc. Am. Mem. 73, 259 pp.

Gall, Q. and Hyde, R. (1989) Analcime in lake and lake-margin sediments of the Carboniferous Rocky Brook Formation, western Newfoundland, Canada. Sedimentology 36, 875-887.

Goble, R.J., Treves, S.B., and Ghazi, A.M. (1993) Comparison of the Rainy ridge analcime phonolite sill and the Crowsnest volcanics, Alberta, Canada. Can. J. Earth Sci. 30, 1644-1649.

Gottardi, G. and Galli, E. (1985) Natural Zeolites. Springer-Verlag, Berlin.

Gottardi, G. and Obradovic, J. (1978) Sedimentary zeolites in Europe. Fortschr. Mineral. 56, 316-366.

Gündogdu, M.N., Yalcin, H., Temel, A., and Clauer, N. (1996) Geological, mineralogical and geochemical characteristics of zeolite deposits associated with borates in the Bigadic, Emet and Kirka Neogene lacustrine basins, western Turkey. Mineral. Deposita 31, 492-513.

Gupta, A.K. and Fyfe, W.S. (1975) Leucite survival: The alteration to analcime. Can. Mineral. 13, 361-363.

Harada, K., Tanaka, K., and Nagashima, K. (1972) New data on the analcime-wairakite series. Am. Mineral. 57, 924-931.

Harker, A. (1954) Petrology for students. Cambridge University Press, Cambridge, England. 283 p.

Haüy, R.-J. (1797) Analcime. J. des Mines 5, 278-279.

Haüy, R.-J. (1801) Traité de minéralogie 3. Chez Louis, Paris, France.

Hay, R.L. (1966) Zeolites and zeolitic reaction in sedimentary rocks. Geol. Soc. Amer., Spec. Pap. 85, 130 pp.

Hay, R.L. (1970) Silicate reactions in three lithofacies of a semi-arid basin, Olduvai Gorge, Tanzania. Min. Soc. Amer., Spec. Pap. 3, 237-255.

Hay, R.L. and Sheppard, R.S. (1977) Zeolites in open hydrologic systems. In: Mumpton, F.A. (ed.) Mineralogy and Geology of Natural Zeolites, Miner. Soc. Am., Short Course Notes, v. 4, 93-102.

Hazen, R.M. and Finger, L.W. (1979) Polyhedral tilting: A common type of pure displacive phase transition and its relationship to analcite at high pressure. Phase Trans. 1, 1-22.

Henderson, C.M.B. and Gibb, F.G.F. (1983) Felsic mineral crystallization trends in differentiating alkaline basic magmas. Contr. Min. Petr. 84, 355-364.

Hey, M.H. (1930) Studies on the zeolites. I. General review. Mineral. Mag. 22, 422-437.

High, L.R. and Picard, M.D. (1965) Sedimentary petrology and origin of analcime-rich Pogo Agie Member, Chugwater Formation (Triassic), west-central Wyoming. J. Sed. Petrol. 35, 49-70.

Horvath, L. and Gault, R.A. (1990) The mineralogy of Mont Saint-Hilaire, Quebec. Mineral. Rec. 21, 284-359.

Iijima, A. (1978) Geological occurrences of zeolite in marine environments. In: Sand, L.B. and Mumpton, F.A. (eds.) Natural Zeolites: Occurrence, Properties, Use, Pergamon Press, Elmsford, New York, 175-198.

Joulia, F., Bonifas, M., Camez, T., Millot, G., and Weid. R. (1959) Découverte d’un important niveau d’analcimolite gresèuse dans le Continental intercalaire de l’ouest de l’Air (Sahara central). Dakar Service de Géologie et de Prospection Minière, 40 pp.

Kastner, M. and Stonecipher, S.A. (1978) Zeolites in pelagic sediments of the Atlantic, Pacific, and Indian Oceans. In: Sand, L.B. and Mumpton, F.A. (eds.) Natural Zeolites: Occurrence, Properties, Use, Pergamon Press, Elmsford, New York, 199-220.

Keith, T.E.C., Thompson, J.M., and Mays, R.E. (1983) Selective concentration of cesium in analcime during hydrothermal alteration, Yellowstone National Park, Wyoming. Geochim. Cosmochim. Acta 47, 795-804.

Keller, W.D. (1952) Analcime in the Popo Agie Member of the Chugwater Formation. J. Sed. Petrol. 22, 70-82.

Keller, W.D. (1953) Analcime in the Chinle formation of Utah correlative with the Pop Agie of Wyoming. J. Sed. Petrol. 23, 10-12.

Koshiya S., Okami, K., Hayasaka, Y., Uzawa, M., Kikuchi, Y., and Doi, N. (1994) On the hydrothermal mineral veins developed in the Takinoue geothermal area, northeast Honshu, Japan. J. Geotherm. Res. Soc. Japan 16, 1-24.

Luhr, J.F. and Giannetti, B. (1987) The Brown Leucitic Tuff of Roccamonfina Volcano (Roman Region, Italy). Cont. Mineral. Petrol. 95, 420-436.

Mazzi, F. and Galli, E. (1978) Is each analcime different? Am. Mineral. 63, 448-460.

Murray, J. and Renard, A.F. (1891) Report on Scientific Results of the Voyage of “H.M.S. Challenger” during the years 1873-1876; Deep Sea Deposits. Johnson Reprint Co., London, 525 pp.

Obradovic, J. (1988) Occurrences and genesis of sedimentary zeolites in Serbia. In: Kalló, D. and Sherry, H.S. (eds.) Occurrence, Properties and Utilization of Natural Zeolites. Akadémiai Kiadó, Budapest, 59-70.

Obradovic, J., Dimitrijevic, R., Visic, N., and Kasanin, M. (1995) Analcime from Teriary lacustrine basins in Serbia. Annales géol. de la péninsula balkanique 59, 255-273 (MA 97M/0839).

Otálora, G. (1964) Zeolites and related minerals in Cretaceous rocks of east-central Puerto Rico. Am. J. Sci. 262, 726-734.

Pearce, T.H. (1970) The analcite-bearing volcanic rocks of the Crowsnest Formation, Alberta. Can. J.   Earth Sci., 7, 46-66.

Pearce, T.H. (1993) Analcime phenocrysts in igneous rocks: Primary or secondary?---Discussion. Am. Mineral. 78, 225-229.

Pekov, I.V. (2000). Lovozero Massif: History, Pegmatites, Minerals. Ocean Pictures Ltd., Moscow, Russia. 484 pp.

Polikarpov, A.I., Polyakovskiy, V. Ya., and Kiseleva, O. V. (1986) Analcime and associated argillaceous minerals in supersaline deposits of the Solikamsk Basin. Zap. Vses. Mineral. Ob. 115, 86-93.

Raade, G., Åmli, R., Mladeck, M.H., Din, V.K., Larsen, A.O., and Åsheim, A. (1983) Chiavennite from syenite pegmatites in the Oslo Region, Norway. Am. Mineral. 68, 628-633.

Remy, R.R. and Ferrell, R.E. (1989) Distribution and origin of analcime in marginal lacustrine mudstones of the Green River Formation, south-central Uintsa Basin, Utah. Clays and Clay Miner. 37, 419-432.

Renaut, R.W. (1993) Zeolitic diagenesis of late Quarternary fluviolacustrine sediments and associated calcrete formation in the Lake Bogoria Basin, Kenya Rift Valley. Sedimentology, 40, 271-301.

Ross, C.S. (1928) Sedimentary analcite. Am. Mineral. 13, 195-197.

Roux, J. and Hamilton, D.L. (1976) Primary igneous analcite--an experimental study. J. Petrol. 17, 244-257.

Saha, P. (1959) Geochemical and X-ray investigation of natural and synthetic analcites. Am. Mineral. 44, 300-313.

Seki, Y. (1969) Facies series in low-grade metamorphism. J. Geol. Soc. Japan 75, 255-266.

Sheppard, R.A. and Gude, A.J. 3rd. (1969) Diagenesis of tuffs in the Barstow Formation, Mud Hills San Bernardino County, California. U.S. Geol. Surv., Prof. Paper  634, 35 pp.

Sheppard, R.A. and Gude, A.J. 3rd. (1973) Zeolites and associated authigenic silicate minerals in tuffaceous rocks of the Big Sandy formation, Mohave County, Arizona. U.S. Geol. Surv., Prof. Paper  830, 36 pp.

Sheppard, R.A. and Hay, R.L. (2001) Formation of zeolites in open hydrologic systems. In: Bish, D.L. and Ming, D.W. (eds.) Natural Zeolites: Occurrence, Properties, Applications, Reviews in Mineralogy & Geochemistry, Mineralogical Society of America, Vol. 45, Washington, D.C., 261-276.

Surdam, R.C. (1966) Analcime-wairakite mineral series. Geol. Soc. Am. Spec. Paper 87, 169-170.

Tang, Z., Parnell, J., and Longstaffe, F.J. (1997) Diagenesis of analcime-bearing reservoir sandstones: the Upper Permian Pingdiquan Formation, Junggar, Basin, northwest China. J. Sediment. Res. 67, 486-498.

Taylor, W.H. (1930) The crystal structure of analcite (NaAlSi2O6•H2O). Z. Kristallogr. 74, 1-19.

Teruggi, M.E. (1964) A new and important occurrence of sedimentary analcime. J. Sed. Petrol. 34, 761-767.

Teruggi, M.E. and Andreis, R.R. (1963) Revision de las zeiktas con especial referencia a su importancia sedimentologia. Revista de la Asoc. Geol. Argentina 18, 73-95.

Utada, M. (1988) Occurrence and genesis of hydrothermal zeolites and related minerals from the Kuroko-type mineralization areas, Japan. In: Kalló, D. and Sherry, H.S. (eds.) Occurrence, Properties and Utilization of Natural Zeolites. Akadémiai Kiadó, Budapest, 39-48.

van de Kamp, P.C. and Leake, B.E. (1996) Petrology, geochemistry, and Na metasomatism of Triassic-Jurassic non-marine clastic sediments in the Newark, Hartford, and Deerfield rift basins, northeastern USA. Chem. Geol. 133, 89-124.

Vanderstappen, R. and Verbeek, T. (1959) Présence d’analcime d’origine sédimentaire dans le Mésozoique du Bassin du Congo. Belge Géol., Paleont., et Hydrol. Bull. 88, 417-421.

Vanderstappen, R. and Verbeek, T. (1964) Analcime et mineraux argileux des formations geologiques de la cuvette congolaise (Republique du Congo). Musee Royal de l’Afrique Centrale, Belgique Annales, Serie in-8°, Sciences Géologiques, no. 47, 88 pp.

Van Houten, F.B. (1960) Composition of Upper Triassic Lockatong argillite, west-central New Jersey. Jour. Geol. 68, 666-669.

Van Houten, F.B. (1962) Cyclic sedimentation and the origin of analcime-rich Upper Triassic Lockatong Formation, west-central New Jersey and adjacent Pennsylvania. Am. Jour. Sci. 260, 561-576.

Van Houten, F.B. (1965) Composition of Triassic Lockatong and associated formations of Newark Group, central New Jersey and adjacent Pennsylvania. Am. Jour. Sci. 263, 825-863.

Whateley, M.K.G., Querol, X., Fernández-Turiel, J.L., and Tuncali, E. (1996) Zeolites in Teriary coal from the Cayirhan mine, Beypazari, Turkey. Mineral. Deposita 31, 529-538.

Wilkinson, J.F.G. (1965) Some feldspars, nephelines and analcimes from the Square Top Intrusion, Nundle, N.S.W. J. Petrol. 6, 420-444.

Wilkinson, J.F.G. (1968) Analcimes from some potassic igneous rocks and aspects of analcime-rich igneous assemblages. Contr. Min. Petr. 18, 252-269.