A geochemical evaluation of perennial spring activity and associated mineral precipitates at...

15
A geochemical evaluation of perennial spring activity and associated mineral precipitates at Expedition Fjord, Axel Heiberg Island, Canadian High Arctic Christopher R. Omelon * , Wayne H. Pollard, Dale T. Andersen Department of Geography, McGill University, 805 Sherbrooke St. W., Montre ´al, Que., Canada H3A-2K6 Received 5 April 2004; accepted 21 August 2005 Editorial handling by I. Hutcheon Available online 15 November 2005 Abstract Two groups of perennial springs are observed in the Canadian High Arctic at Expedition Fjord on Axel Heiberg Island at Colour Peak and Gypsum Hill. Saline discharge (1.3–2.5 molal NaCl) produces a variety of calcite (travertine) and gypsum-rich precipitates. Saturation index calculations of the spring waters at Colour Peak suggest CO 2 degassing from the waters causes calcite precipitation. Gypsum precipitation dominates at Gypsum Hill, where spring waters have lower alkalinity and higher SO 4 concentrations. Mineral accumulations form both channel and rimstone pool morphologies as a result of varying slope conditions. At Colour Peak, confined flow in steep slope areas develop massive structures in con- trast to more friable, porous accumulations in areas where waters fan out on shallower slopes; these morphological vari- ations lead to corresponding varying apparent rates of mineral precipitation. Mineral precipitation at Gypsum Hill is far less notable as a result of lower discharge rates and annual degradation by icing formation. Microscopic observations and geochemical analyses of the channel precipitates at Colour Peak reveal alternating light (calcite spar) and dark (anhedral microcrystalline calcite combined with organic matter and non-carbonate minerals) laminae. Rimstone pools forming in lower sections of spring discharge are composed of accumulations of large euhedral calcite crystals interbedded with allo- chthonous inputs. High concentration of dissolved solids is responsible for slow travertine precipitation rates, which occurs during winter. This precipitation is further retarded during summer months by the introduction of crystal growth inhib- itors such as Fe 3+ and deposition of organic matter and soil sediments. Ó 2005 Elsevier Ltd. All rights reserved. 1. Introduction Travertine-depositing springs are natural systems that offer a unique opportunity to study the various mechanisms by which waters establish equilibrium with conditions at the earthÕs surface (Amundson and Kelly, 1987). Over the past several decades, interest in these environments has been initiated from a number of scientific fields: for geochemists, travertine-precipitating waters provide an opportu- nity to evaluate a dynamic carbonate system and the major controls associated with calcite precipita- tion (Barnes, 1965; Jacobson and Usdowski, 1975; 0883-2927/$ - see front matter Ó 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.apgeochem.2005.08.004 * Corresponding author. Present address: Department of Geology, University of Toronto, 22 Russell Street, Toronto, Ontario, Canada M5S-3B1. Tel.: +1 416 978 0826; fax: +1 416 978 3938. E-mail address: [email protected] (C.R. Omelon). Applied Geochemistry 21 (2006) 1–15 www.elsevier.com/locate/apgeochem Applied Geochemistry

Transcript of A geochemical evaluation of perennial spring activity and associated mineral precipitates at...

Applied Geochemistry 21 (2006) 1–15

www.elsevier.com/locate/apgeochem

AppliedGeochemistry

A geochemical evaluation of perennial spring activityand associated mineral precipitates at Expedition Fjord, Axel

Heiberg Island, Canadian High Arctic

Christopher R. Omelon *, Wayne H. Pollard, Dale T. Andersen

Department of Geography, McGill University, 805 Sherbrooke St. W., Montreal, Que., Canada H3A-2K6

Received 5 April 2004; accepted 21 August 2005Editorial handling by I. HutcheonAvailable online 15 November 2005

Abstract

Two groups of perennial springs are observed in the Canadian High Arctic at Expedition Fjord on Axel Heiberg Islandat Colour Peak and Gypsum Hill. Saline discharge (�1.3–2.5 molal NaCl) produces a variety of calcite (travertine) andgypsum-rich precipitates. Saturation index calculations of the spring waters at Colour Peak suggest CO2 degassing fromthe waters causes calcite precipitation. Gypsum precipitation dominates at Gypsum Hill, where spring waters have loweralkalinity and higher SO4 concentrations. Mineral accumulations form both channel and rimstone pool morphologies as aresult of varying slope conditions. At Colour Peak, confined flow in steep slope areas develop massive structures in con-trast to more friable, porous accumulations in areas where waters fan out on shallower slopes; these morphological vari-ations lead to corresponding varying apparent rates of mineral precipitation. Mineral precipitation at Gypsum Hill is farless notable as a result of lower discharge rates and annual degradation by icing formation. Microscopic observations andgeochemical analyses of the channel precipitates at Colour Peak reveal alternating light (calcite spar) and dark (anhedralmicrocrystalline calcite combined with organic matter and non-carbonate minerals) laminae. Rimstone pools forming inlower sections of spring discharge are composed of accumulations of large euhedral calcite crystals interbedded with allo-chthonous inputs. High concentration of dissolved solids is responsible for slow travertine precipitation rates, which occursduring winter. This precipitation is further retarded during summer months by the introduction of crystal growth inhib-itors such as Fe3+ and deposition of organic matter and soil sediments.� 2005 Elsevier Ltd. All rights reserved.

1. Introduction

Travertine-depositing springs are natural systemsthat offer a unique opportunity to study the various

0883-2927/$ - see front matter � 2005 Elsevier Ltd. All rights reserved

doi:10.1016/j.apgeochem.2005.08.004

* Corresponding author. Present address: Department ofGeology, University of Toronto, 22 Russell Street, Toronto,Ontario, Canada M5S-3B1. Tel.: +1 416 978 0826; fax: +1 416978 3938.

E-mail address: [email protected] (C.R. Omelon).

mechanisms by which waters establish equilibriumwith conditions at the earth�s surface (Amundsonand Kelly, 1987). Over the past several decades,interest in these environments has been initiatedfrom a number of scientific fields: for geochemists,travertine-precipitating waters provide an opportu-nity to evaluate a dynamic carbonate system andthe major controls associated with calcite precipita-tion (Barnes, 1965; Jacobson and Usdowski, 1975;

.

2 C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15

Herman and Lorah, 1987; Pentecost, 1992, 1995).This work has commonly progressed in conjunctionwith studies examining isotope fractionation associ-ated with natural calcite precipitation from springwaters (Friedman, 1970; Usdowski et al., 1979;Dandurand et al., 1982; Amundson and Kelly,1987; Usdowski and Hoefs, 1990; Hamilton et al.,1991; Liu et al., 2003) as well as the kinetics (Lorahand Herman, 1988) and associated deposition rates(Herman and Lorah, 1988; Dreybrodt et al., 1992;Liu et al., 1995; Bono et al., 2001) of CaCO3 in trav-ertine environments.

The role of plants (Barnes, 1965), algae andmosses (Emeis et al., 1987; Weijermars et al., 1986;Hoffer-French and Herman, 1989; Pentecost,1996), and bacteria (Chafetz and Folk, 1984; Folk,1993; Pentecost et al., 1997) in travertine precipita-tion has created a new impetus to study biomineral-ization not only in travertine environments but in allcarbonate sediments (Riding, 2000). The mineralog-ical and fabric variability of travertine has also re-ceived significant attention from sedimentologistsattempting to determine both environmental (Loveand Chafetz, 1988; Chafetz et al., 1991a; Guo andRiding, 1992; Jones and Renaut, 1996) and biolog-ical (Folk et al., 1985; Folk, 1993; Guo and Riding,1994; Fouke et al., 2000) influences on travertinegrowth and diagenesis. The presence of laminae intravertine on daily (Chafetz and Folk, 1984; Folket al., 1985; Guo and Riding, 1992) and seasonal(Chafetz et al., 1991b; Matsuoka et al., 2001; Kanoet al., 2003) cycles reflecting biological activity and/or physico-chemical changes continues to be aninteresting theme in travertine studies.

The springs at Expedition Fjord, Axel HeibergIsland in the Canadian High Arctic display interest-ing combinations of water chemistries and associ-ated mineral precipitates that are stronglyinfluenced by the polar climate of this region. Thesprings were first reported by Beschel (1963) andhave received subsequent interest in recent years(Pollard et al., 1998, 1999; Omelon et al., 2001;Andersen et al., 2002), and continue to be a focusof research in hydrology, geochemistry, sedimentol-ogy and microbiology. The emergence of these‘‘cold spring’’ waters (Pentecost et al., 2003) andtheir re-equilibration with the atmosphere causesthe precipitation of ‘‘saline travertine’’ (cf. Fordand Pedley, 1996). In light of discussions on termi-nology, these descriptions are used with some cau-tion. Many aspects of the mineral precipitatesappear similar to those of more temperate-climate

travertine springs, however clear differences betweenthese springs and other natural spring systems sug-gests that the Expedition Fjord springs be consid-ered one end-member to the global spectrum oftravertine-depositing springs.

Studying brines associated with sedimentary ba-sins (e.g., Tijani, 2004) and more specifically, brinesrelated to petroleum reservoirs (e.g., Losh et al.,2002) is important for understanding water–rockinteractions and the formation and migration ofthese reservoir fluids, as well as hydrocarbon matu-ration related to rising salt structures such as thosepresently found in the Gulf of Mexico. Extensiveexploration for oil and gas in the Canadian Arcticin the 1970s suggests that the brine chemistry ofthe spring waters at Expedition Fiord are relatedto underground reservoirs, but the authors are una-ware of published studies using them as indicatorsfor hydrocarbons. In addition, evidence of perennialdischarge in continuous permafrost environmentsmay provide direct evidence of thermal anomaliesassociated with the salt diapirs that can play animportant role in the development of hydrocar-bons at shallower depths than those encounteredin non-polar regions (M. Zentilli, personal commu-nication). The objective of this paper is to betterdocument the physical and chemical characteristicsof perennial spring discharge at Expedition Fjordand relate these conditions to associated mineralprecipitates in light of the dominating polar condi-tions of this region.

2. Study site

The two study areas are located on Axel HeibergIsland (79�26 0N, 90�46 0W) approximately 11 kmapart along the northern boundary of ExpeditionFjord and the associated Expedition River flood-plain (Fig. 1(a)). Perennial springs emerge fromthe base of gypsum-anhydrite diapirs (piercementdomes) composed of Upper Mississippian to MiddlePennsylvanian evaporites that were originallydeposited as a thick (>400 m) bedded sequence(Otto Fjord Formation; Nassichuk and Davies,1980) within the Sverdrup Basin, an area which cov-ers no more than one-third of the Canadian ArcticArchipelago. The diapiric structures formed byintrusion through the overlying sediments duringTertiary mountain building (Hoen, 1964). Of the>100 diapirs of the Sverdrup Basin (Thorsteinsson,1974), perennial spring activity has only been con-firmed at the two study localities.

Fig. 1. (a) Axel Heiberg Island study area and Expedition Fjord. Arial photographs of perennial spring activity at (b) Colour Peak and(c) Gypsum Hill.

C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15 3

The region encompassing Expedition Fjord ischaracterized by polar desert conditions with cold,drywinters and cool summers. An automatic weatherstation approximately 1 km from Gypsum Hill and11 km from Colour Peak indicates a mean annualair temperature of�15 �C (Doran et al., 1996).Meanmonthly air temperatures for Eureka, locatedapproximately 120 km NE of Expedition Fjord are�36.1 �C and 5.4 �C for January and July, respec-tively (Pollard et al., 1999).

Axel Heiberg Island lies within the zone of con-tinuous permafrost, defined as any rock or soilmaterial that has remained below 0 �C continuouslyfor two or more years (Brown, 1972). Permafrostdepth has not been measured at Expedition Fjord,but a permafrost thickness of >400 m was docu-mented in an exploration well near Mokka Fjordon the east side of Axel Heiberg Island roughly60 km from the study area. Other wells in the area

indicate that permafrost is generally between 400and 600 m deep (Taylor and Judge, 1976).

Approximately 20 outlets over a 900 m2 areamake up the springs at Colour Peak (Fig. 1(b)). Dis-charge rates range from 1.3 to 1.8 L s�1 (Pollardet al., 1999) with an estimated total discharge of20–25 L s�1. The spring outlets are stationary, lead-ing to the development of dominant discharge flowpaths along which mineral precipitates accumulate.In contrast, groundwater activity at Gypsum Hillconstitutes large zones of saturated soil interspersedwith approximately 40 springs and seeps over a sim-ilar sized area (Fig. 1(c)). Average discharge ratesbetween individual springs are highly variable, butquantitative measurements taken from specific out-lets remain constant; the largest spring outlet at thissite has a discharge rate between 0.9 and 1.0 L s�1.Total discharge at Gypsum Hill is estimated to beapproximately 10–15 L s�1 (Pollard et al., 1999).

4 C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15

3. Methodology

3.1. Field methods and sample collection

Water and mineral samples were collected duringfield trips in July 1997, April–May 1998, and July1998 and returned to McGill University for analy-ses. In addition to general sampling of spring dis-charge and mineral precipitates throughout theareas, 3 spring systems were chosen for detailed,long-term chemical monitoring (Fig. 2) due to thefact that these springs display the largest accumula-tions of precipitate. These included two at ColourPeak (named hereafter as CP-1 and CP-2) and oneat Gypsum Hill (GH-1). Measurements for pHand temperature were taken during each visit tothese sites; specific conductivity, salinity, and totalalkalinity were measured in the field laboratory.

During each field trip, water samples for anionand cation analyses were collected in acid-washed125 and 250 ml polyethylene bottles. Samples forlaboratory alkalinity titrations were collected inacid-washed 125 ml dissolved oxygen bottles withground glass stoppers; titrations were performed

Fig. 2. Profile sketches for springs CP-1 and CP-2

within 1–2 h after collection by the addition of0.01 M H2SO4 to a 50 ml aliquot. Samples collectedfor H2S analysis were fixed by the addition of 200 ll1.0 M zinc acetate solution [Zn(C2H3O2)2 Æ 2H2O]per 100 ml sample at collection time.

Mineral samples were collected at selected pointsfrom CP-1, CP-2 and GH-1, representing all the ma-jor precipitates found at these sites. The sampleswere individually placed in sterile roll-top bags.

3.2. Laboratory methods

All water samples for cation and anion analyseswere filtered through 0.45 lm Millipore filters uponreturn from the field. Samples for cation analyseswere acidified with 0.5 ml concentrated (16 M)HNO3 per 100 ml sample within days of collection.These were analyzed for Ca, Mg, Na and K byinductively coupled plasma atomic emission spec-troscopy. Unacidified samples were analyzed forSO4, Cl, NO3 and PO4 by ion chromatography. Inboth cases, multiple dilutions were required to bringelement concentrations within measurable ranges;all samples were run in triplicate and were found

at Colour Peak and GH-1 at Gypsum Hill.

C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15 5

to range by no more than 10%. Hydrogen sulphideconcentrations were determined spectrophotometri-cally following the procedure of Cline (1969).

Bulk precipitate mineralogy was determined bypowdering the samples, which were then mixed withacetone and spread over glass slides and analyzedusing a Philips PW 1710 X-ray diffractometer. Sam-ples for scanning electron microscopy were carbon-coated and analyzed using a JEOL 8900 X-raymicroanalyzer equipped with energy dispersive (EDS)and wavelength dispersive (WDS) spectrometers.

3.3. Data analysis

Chemical concentration data and field pH, alka-linity and temperature measurements for each sitewere input into PHRQPITZ (Plummer et al.,1988), a thermodynamic data-based computer pro-gram for geochemical calculations of brines andother high ionic strength electrolyte solutions. Theprogram uses the Pitzer virial-coefficient approachfor activity-coefficient corrections.

The program calculates the saturation state ofthe water with respect to solid mineral phases bythe equation:

SIx ¼ logIAPðT ÞKxðT Þ

� �; ð1Þ

where SIx is the saturation index of mineral x,IAP(T) is the ion activity product at specified tem-perature T (�C), and Kx is the equilibrium solubilityproduct of mineral x. The resulting values indicatewhether the solution is undersaturated (�SIx),supersaturated (+SIx), or at equilibrium (SIx = 0)with respect to the mineral in question. The solu-tions modeled were charge balanced by adding Cl�.

In order to explore how disequilibrium conditionsexist in the spring waters under the large climate vari-ations observed and measured throughout the year(Doran et al., 1996), each water chemistry data set(including pH and alkalinity) was modeled at 5 �Ctemperature intervals from�5 to +20 �C. The resultsare presented as the various hypothetical saturationstates of the dominant minerals in the springs overthis temperature range (see below).

Theoretical CO2 partial pressures ðPCO2Þ were

calculated by the equation:

log PCO2¼ �pHþ log

cHCO3mHCO�

3

� �KCO2

K1

; ð2Þ

where KCO2is the equilibrium (Henry�s law) con-

stant for CO2 and K1 is the first dissociation con-

stant of H2CO3 (Langmuir, 1997, p. 158). Thisassumes that HCO�

3 is the main contributor to alka-linity, which is the case for much of the spring waterprofiles. These values correspond well to direct mea-surements taken at the springs.

4. Results

4.1. Water chemistry

4.1.1. Chemical data and the carbonate system

Chemical analyses for spring discharge watersat Colour Peak and Gypsum Hill are presented inTable 1. The high specific conductivities (120–230 mS cm�1) of the waters are attributed to highconcentrations of Na and Cl ions as well as signifi-cant inputs of Ca and SO4. [Cl]/[Na] (molal) for thetwo springs closely compare with that of seawater,ranging from 0.84 to 1.03, and [Ca]/[SO4] valuesrange from 3.78 to 1.36 (Colour Peak and GypsumHill, respectively). Dissolved sulphide concentra-tions range from 400 to 412 lmol L�1 at the springoutlets.

Slope profiles and spring and summer measure-ments of temperature, pH, and alkalinity alongthe discharge paths of CP-1, CP-2 and GH-1 areshown in Fig. 3. Pollard et al. (1998) show thatthe highest spring discharge temperatures remainconstant (�6.5 �C) throughout the year, even whenair temperatures reach lows near �40 �C; similarconsistency in discharge temperatures is measuredin the springs studied here, but there is considerablevariability in temperatures away from the outletsboth between and within seasons. During springmonths, low air temperatures effectively cool thewaters after they emerge, whereas warm air temper-atures and highly variable insolation rates duringsummer months lead to varying degrees of warmingof the waters as they flow downslope. Spring outletpH values range from slightly acidic at Colour Peakto circum-neutral at Gypsum Hill, and consistentlyincrease with distance away from the outlet withmaximum values of 9.5 and 9.0 at Colour Peakand Gypsum Hill, respectively. In contrast to watertemperatures, rates of pH change over distance aresimilar in both spring and summer. Alkalinities weremeasured only in spring, but in all cases these showa general decrease with distance.

4.1.2. Geochemical modeling

Input data and resulting output calculations foreach site at temperatures of �5, +5 and +20 �C

Table 1Major ion chemistry of waters collected from spring outlets at Colour Peak and Gypsum Hill

Colour Peak spring Gypsum Hill spring

Spring 1 Spring 2

Temperature (�C) 6.2 6.4 6.6pH 6.91 6.82 7.43Conductivitya (mS/cm) 227 230 120Alkalinityb 181.5 183.5 62.6PCO2

(bar) 10�1.92 10�1.82 10�2.89

Density (g/cm3) 1.114 1.056Ca2+ 3628 2115Mg2+ 316 115Na+ 52900 27927K+ 186 52.5Cl� 68600 44212SO2�

4 2300 3724PO3�

4 LDL LDLNO�

3 0.001 LDLHS� (lmol/l) 412.0 400.0DICc 17.2 13.1

Results expressed in mg/L unless otherwise indicated.LDL: Na, Mg: 0.01 mg/L; Ca: 0.10 mg/L; K: 0.05 mg/L; Cl, SO4, PO4, NO3: 1 mg/L.a Measured at 25 �C.b Alkalinity expressed as total CaCO3 (mg/L) at 25 �C.c DIC expressed as carbon.

Colour Peak CP-1 Colour Peak CP-2 Gypsum Hill GH-1

-5

0

5

10

15

20

tem

per

atu

re (

oC

)

6.5

7.0

7.5

8.0

8.5

9.0

9.5

10.0

pH

120

130

140

150

160

170

180

190

0 5 10 15 20 25 30 35 40

Distance from outlet (m)

mg

. L-1

CaC

O3

-5

0

5

10

15

20

6.5

7.0

7.5

8.0

8.5

9.0

9.5

10.0

120

130

140

150

160

170

180

190

0 15 30 45 60 75 90 105 120 135 150

Distance from outlet (m)

45

50

55

60

hei

gh

t a.

s.l.

(m)

channels, laminate precipitates

rimstone pools,euhedral calcite crystals

0

10

20

30

40

50

60

70channels, tunnels, cascades,

laminate precipitatesrimstone pools,

euhedral calcite crystals

-5

0

5

10

15

20

6.5

7.0

7.5

8.0

8.5

9.0

9.5

10.0

303540455055606570

0 10 20 30 40 50

Distance from outlet (m)

0

5

10

15

20channels, pools, small cascades, gypsum-rich laminate precipitates

gypsum-rich barrage pools

springsummer

springsummer

springsummer

a

b

c

d

Fig. 3. Profiles of important parameters illustrating the downstream evolution of spring chemistry at Colour Peak and Gypsum Hill.(a) simplified geomorphological profile displaying sample sites for geochemical modeling; (b) changes in temperature, (c) pH, and(d) alkalinity with distance. Error bars represent the maximum range of multiple measurements.

6 C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15

are presented in Table 2. In addition, calculated sat-uration indices for calcite (SIC) and gypsum (SIG)and associated changes in calculated PCO2

values

for CP-1, CP-2 and GH-1 are shown graphicallyin Fig. 4 where saturation indices for each mineralare presented for the entire modeled temperature

C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15 7

range (�5 to +20 �C), listed in Table 3. Spring out-let waters at both CP-1 and CP-2 are supersaturatedwith respect to calcite over all modeled tempera-tures, ranging from 0.200 to 0.596 at CP-1 and0.115 to 0.513 at CP-2. All but the low-end temper-ature measurement (�5 �C) is above calcite satura-tion at the GH-1 outlet (�0.0991). Althoughtemperature does influence calcite solubility(Fig. 4), the largest increases in SIC values occuras the waters move downslope – concomitant withdecreasing PCO2

– reaching highs of 2.061 and2.084 for CP-1 and CP-2 and 1.328 for GH-1 (at+20 �C). In contrast, the waters are undersaturatedwith respect to gypsum for all points and tempera-tures modeled at CP-1 and CP-2, with SIG valuesranging from �0.099 to �0.152 and �0.099 to�0.152, respectfully. Gypsum saturation conditionsare higher at GH-1, ranging from slightly oversatu-rated at all points at �5 �C (0.020–0.019 betweenthe outlet and 44 m downslope) to slightly undersat-urated (�0.038 to �0.039) at +20 �C. Unlike calcite,modeling shows that gypsum solubility conditionsremain relatively constant throughout the dischargeprofile and are influenced more by changes in tem-perature than PCO2

values.

Colour Peak CP-1 Colour Pea

-0.5

0.0

0.5

1.0

1.5

2.0

2.5

0 50

Distance from

CaCO3

Ca

-6

-5

-4

-3

-2

-1

0

-6

-5

-4

-3

-2

-1

0

log

PC

O2

-0.5

0.0

0.5

1.0

1.5

2.0

2.5

0 10 20 30 40

Distance from outlet (m)

SI CaCO3

CaSO4.2H2O

a

b

Fig. 4. Theoretical PCO2concentrations and saturation indices (SI) fo

relationship between decreasing PCO2concentrations and increasing

carbonate chemistry as shown by constant SI values throughout the enrange of multiple measurements.

Table 2Input chemical data for waters at selected points along spring flow pat

Element (molality) Ca Mg

CP-1, CP-2 9.18 · 10�2 1.32 · 10�2

GH-1 5.38 · 10�2 4.82 · 10�3

4.2. Mineral precipitates

4.2.1. Colour peak

Well-developed channels, tunnels and cascadestructures ranging from millimetres to tens of cen-timetres in thickness characterize the dominantprecipitates in the upper sections of spring over-flow at CP-1 and CP-2 (Fig. 5). The channelsare generally U-shaped with steep sides androunded bottoms. In lower sections, lobate depos-its form friable, loosely cemented rimstone poolsand mounds as a result of a break in slope(Fig. 6(a)). Rimstone pools associated with dis-charge from CP-1 range in size from 1 to 10 cmlaterally and are 0.2–0.5 cm deep, composed ofeuhedral crystals 0.25–0.5 cm in length(Fig. 6(b)). Smaller rimstone pools made up ofeuhedral–subhedral crystals averaging 0.1–0.25 cmlong characterize the lower region of spring over-flow at CP-2. Total mineral accumulations in bothareas are normally 10–15 cm thick, and wind-deposited organic matter and sediment progres-sively accumulates in the bottom of the poolsthroughout the summer season. Bulk X-ray dif-fraction analyses of all mineral precipitates reveal

k CP-2 Gypsum Hill GH-1

-0.5

0.0

0.5

1.0

1.5

0 10 20 30 40

Distance from outlet (m)

CaCO3

CaSO4.2H2O

-6

-5

-4

-3

-2

-1

0

100 150

outlet (m)

SO4.2H2O

r spring profiles at Colour Peak and Gypsum Hill. Note inverseSI for calcite. Gypsum solubility is independent of changes intire spring discharge profile. Error bars represent the maximum

hways at CP-1, CP-2 and GH-1

Na K Cl S

2.33 4.82 · 10�3 1.96 2.43 · 10�2

1.24 1.63 · 10�3 1.27 3.95 · 10�2

Tab

le3

CalculatedoutputdataforpH,totalalkalinity,

molesofClad

ded

andsaturationindices

forcalcite,gypsum

andPCO

2forwatersat

selected

pointsalongspringflowpathway

sat

CP-1,

CP-2

andGH-1

attemperaturesof�5,

5an

d20

�C

Distance

from

outlet

(m)

pH

Total

alkalinity(M

)Molesof

Clad

ded

SIcalcite

SIgy

psum

SIPCO

2

�5�C

5�C

20�C

�5�C

5�C

20�C

�5�C

5�C

20�C

CP-1

06.91

3.67

870.53

390.19

950.37

290.59

60�0.09

87�0.13

01�0.14

79�1.97

19�1.91

78�1.81

798

7.26

3.58

750.53

40.53

310.70

440.92

30�0.09

89�0.13

03�0.14

81�2.33

82�2.28

61�2.19

0736

9.38

2.81

110.53

471.95

302.01

562.06

08�0.10

26�0.13

41�0.15

22�5.15

43�5.21

09�5.28

86

CP-2

06.82

3.71

930.53

380.11

510.28

880.51

27�0.09

86�0.13

�0.14

78�1.87

63�1.82

19�1.72

1335

7.66

3.62

190.53

390.99

271.08

861.29

58�0.09

91�0.13

05�0.14

85�2.74

84�2.70

16�2.61

7463

8.33

3.15

240.53

41.49

981.64

031.79

88�0.10

00�0.13

17�0.15

01�3.51

02�3.48

66�3.45

2811

08.57

3.44

560.53

411.66

831.79

251.92

35�0.10

07�0.13

24�0.15

09�3.82

11�3.81

58�3.80

7314

79.39

2.95

30.53

461.97

682.03

892.08

35�0.10

27�0.13

43�0.15

24�5.15

03�5.20

74�5.28

57

GH-1

07.43

1.27

420.00

7609

�0.09

910.07

470.29

810.02

030.01

49�0.03

80�2.92

94�2.88

47�2.79

9428

8.55

0.95

255

0.00

793

0.81

470.96

071.12

920.01

97�0.01

56�0.03

89�4.25

50�4.23

8�4.20

7344

8.92

0.93

627

0.00

7946

1.08

241.20

271.32

780.01

94�0.01

6�0.03

93�4.72

69�4.73

56�4.74

83

8 C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15

a composition of CaCO3 as calcite. Evaporitemineral accumulations of halite (NaCl) and gyp-sum (CaSO4 Æ 2H2O) are observed both in ground-water–saturated soils around the spring outletsand in splash zones proximal to the spring waters.

Inspection of mineral precipitates forming up-stream channels in polished thin sections revealscomplex laminae (Fig. 7) consisting of alternatinglayers of white sparite crystals and darker micritecrystals. Aggregates of these layers form concentriclaminae radiating from nucleation points, giving thebulk precipitate surface a mammillary appearancethat is black in colour. The white layers are com-posed of calcite, exhibiting characteristic 60–120�angled cleavage and extreme birefringence underpolarized light. EDS confirmed calcite identifica-tion. Calcite precipitates as acicular, jagged euhe-dral crystals averaging 100–150 lm in thicknesswith no apparent organic component. Crystalsnucleating on the top of the previous layer grow up-wards until an abrupt transition into the dark micri-tic layer. These darker laminae are composed ofanhedral calcite crystals averaging 1–5 lm in diam-eter with no clear growth orientation. The crystalsare coated with an orange to reddish-brown filmand are interspersed with organic matter and soilparticles. These laminae average 5–10 lm thick witha maximum observed thickness of 20 lm. WDSmapping over a 400 lm2 area spanning across thisregion shows accumulations of Fe, S, SiO2, Al,and Na (Fig. 8). High silica concentrations are ob-served mainly in ‘‘valleys’’ or depressions of micriticlaminae and are interpreted as accumulations ofsediment particles. Iron, Na and S concentrationsare distributed more uniformly across the micritelayer and decrease in relative concentrations. Cal-cium concentrations are notably lower within themicrite laminae, but are clearly dominant withinthe calcite spar.

Detailed examination of the rimstone poolsformed in the lower reaches of the spring systemsare characterized by alternating layers of thick (1–2 cm) porous mineral accumulations composed oflarge euhedral calcite crystals and 0.5 mm thickhorizons containing allochthonous organic matterand soil sediment.

4.2.2. Gypsum Hill

Mineral accumulations associated with springdischarge at GH-1 are limited to channels nearspring outlets and small terraced pools furtherdownstream. Based on XRD analysis, gypsum is

Fig. 6. (a) Rimstone pools (terraced mounds) forming in lower sections of spring discharge at Colour Peak. Backpack for scale. (b) Close-up of calcite crystals, which can reach 0.5 cm in length. Allochthonous organic matter and sediment debris accumulates in these areasduring summer months. Scale bar = 2 cm.

Fig. 5. Mineral precipitates associated with spring discharge at Colour Peak. (a) Cascade, (b) channel, and (c) tunnel morphologies.

C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15 9

the only precipitate forming in direct contact withthe spring waters. Halite and accumulations of na-tive S are observed on bare sediment surfaces near

Fig. 7. Photomicrographs under plain polarized light of polished thin slayers are calcite spar, interbedded with microcrystalline calcite layers. Sas nucleation sites for calcite precipitation. Scale bar = 2.5 mm.

the springs, but not directly within the discharginggroundwaters. Precipitated gypsum from the springwaters have much smaller maximum thicknesses

ections of channel travertine from Colour Peak spring. (a) Whitecale bar = 5 mm. (b) Sediment debris within the travertine acting

Fig. 8. WDS map of channel precipitate from Colour Peak. Note spatial distribution of elements: silica predominates in valleys (asallochthonous sediment debris) whereas Fe, Na and S are found (in varying concentrations expressed as intensity) throughout the micritelayer. Calcium (as calcite) is the dominant element in the spar layers. Scale bar 500 = lm.

10 C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15

compared to the calcite precipitating from thesprings at Colour Peak (1–5 cm as opposed to10–25 cm), however it still forms massive accumula-tions composed of dendritic-shaped crystals. Lami-nations occasionally found in channel regions nearthe spring outlets consist of 100–200 lm thick gyp-sum horizons interspersed by thin accumulationsof allochthonous leaf litter and sediment debris.

5. Discussion

5.1. Water chemistry

The high dissolved solid concentrations and [Cl]/[Na] suggests that the salinity of these springs has anorigin related to evaporite dissolution. Observedhigh [Ca]/[Mg] is normally characteristic of karst

C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15 11

aquifers with low dolomite content, but [Ca]/[SO4]values suggests that dissolution of gypsum and/oranhydrite contributes significantly to Ca inputs.Much higher [Ca]/[SO4] values in the waters at Col-our Peak are likely due to the presence of largeramounts of calcite within the diapir. This wouldcorrespond well with the fact that the springsemerge from anhydrite diapirs, which are composedof gypsum and anhydrite in addition to smalleramounts of calcite (Hoen, 1964). Fracture systemsassociated with diapirism or the diapirs themselvesare likely the conduit geology through which thisgroundwater reaches the surface, and may be thesource for this water–rock interaction.

Considering the close proximity of the springs atColour Peak and Gypsum Hill to each other, it isnot unreasonable to suggest similar hydrogeologicalpathways for these two systems as proposed byAndersen et al. (2002). However, differences inwater chemistry do exist, and explanations for thesedifferences (most notably lower alkalinity concen-trations and higher SO4 at Gypsum Hill) remainunresolved. Possible scenarios include: (1) diapirformations with different mineralogical composi-tions, (2) lower activity rates of biological SO4

reduction at Gypsum Hill, or (3) different ground-water pathways resulting from an impermeable bar-rier between the two sites.

5.2. Carbonate system

Alkalinity values in the spring waters at ColourPeak and Gypsum Hill are lower than those re-ported for other travertine-depositing cold-watersprings. Moderate concentrations of dissolved sul-phide in the waters suggests that SO4 reduction isa major contributor of HCO3 alkalinity to the sys-tem by the reaction:

SO2�4 þ 2CH2O ! HS� þ 2HCO�

3 þHþ ð3Þassociated with SO4-reducing bacteria.

Saline and highly brackish spring waters rich insulphide are reported in Poland (Zuber et al.,1997), Spitzbergen (Lauritzen and Bottrell, 1994),and in the Canadian Mackenzie Valley area (vanEverdingen et al., 1982), but only S-rich springs inYellowstone National Park (Mammoth HotSprings, Fouke et al., 2000), Harrogate (UnitedKingdom, Bottrell et al., 1996) and northern Elles-mere Island (Grasby, 2003) report CaCO3-mineralprecipitation in association with spring discharge.Calcite saturation in the waters is likely the result

of a common ion effect caused by dissolution ofgypsum (CaSO4 Æ 2H2O). As the waters move down-slope, degassing of dissolved CO2 to attain equilib-rium with atmospheric conditions is observed asincreasing pH values, decreasing alkalinity and logPCO2

values (Figs. 3 and 4), and increasing calcitesaturation conditions (Fig. 4) by the reversereaction:

Ca2þ þ 2HCO�3 $ CaCO3 ðsÞ þH2Oþ CO2 ðgÞ "

ð4ÞObserved differences in mineral accumulation vol-umes reflecting precipitation rates at CP-1 andCP-2 correlate to morphological variability: flowconfined in channels in the upper sections limitair–water exchange compared with higher rates ofCO2 degassing and higher calcite precipitation ratesin the lower terraced mound sections.

The dominance of gypsum precipitation at GH-1over calcite is attributed to both the dissolved ionchemistry and alkalinity concentrations. HigherSIG at GH-1 (Fig. 4) due to higher SO4 concentra-tions correspond to the observed accumulations ofgypsum that form channels and terraced poolsalong the spring discharge pathways. In addition,the waters have lower SIC values than CP-1 andCP-2 throughout the spring profile due to lowerCa concentrations and alkalinity values.

5.3. Mineral precipitation

Model calculations predicting supersaturatedconditions and expected degree of precipitation ofcalcite contrasts with observed low total mass oftravertine accumulation, which may be due to lowalkalinity concentrations or kinetic inhibitory pro-cesses. Previous studies of calcite-precipitatingsprings show that calcite does not precipitate instan-taneously as supersaturation is reached; in somecases, 10 times supersaturation is attained beforecalcite precipitation is observed (Jacobson andUsdowski, 1975; Dandurand et al., 1982; Lorahand Herman, 1988). There are several reasons to ex-plain this kinetic inhibition, including (1) the natu-ral energy barrier to calcite nucleation (Berner,1981) or a lack of suitable nucleation sites; alterna-tively, high concentrations of dissolved impuritiesmay influence the precipitation mechanism by (2)blocking surface sites for crystal growth, or (3)decreasing precipitation reaction rates (Zuddasand Mucci, 1994). Although Mg2+ has been shownto inhibit calcite growth by substitution for Ca2+

12 C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15

into the crystal lattice (Reddy and Nancollas, 1976;Morse et al., 1979; Morse, 1983) the low [Mg]/[Ca]in these waters (0.14 for Colour Peak, 0.09 for Gyp-sum Hill) would suggest that inhibition by Mg2+

does not play a major role in these spring waters.Orthophosphates can act as strong growth inhibi-tors by blocking active surface sites on the calcitecrystal (Reddy, 1977; Morse, 1983), but measured½PO3�

4 � < 1 mg L�1 suggests a situation similar toMg2+. In addition, growth rate reduction experi-ments by Meyer (1984) and Katz et al. (1993) showthat several other ions have a strong inhibitory effecton calcite growth, including Fe2+ (down to<10�9 M) and Fe3+, which is present in the thinly-bedded micrite laminae in mineral precipitates atCP-1 and CP-2 (Fig. 8). Other solutes such as humicand fulvic acids and organic phosphates may beimportant kinetic inhibitors (Reddy, 1983; Morse,1983; Mitterer and Cunningham, 1985). It is be-lieved that the combined effects of (1) high dissolvedsolid concentrations and the presence of (2) Fe and(3) allochthonous debris all play a role in inhibitingrapid, perennial calcite precipitation from thewaters.

Multi-year accumulations of mineral precipitateat Gypsum Hill are much lower than at ColourPeak, a consequence of observed low rates of springdischarge and associated mineral precipitation. Inaddition, spring discharge leads to the developmentof large icings (Pollard et al., 1998) that form duringwinter months and cover much of the local surface,promoting physical destruction and chemical disso-lution of the mineral accumulations when the icingsmelt in spring.

5.3.1. LaminationsAlthough the development of laminations in trav-

ertine accumulations is often reported in the litera-ture, the incorporation of Fe and other trace metalions into carbonate precipitates in spring environ-ments is not common. High concentrations of Feand Mn (as Fe2O3 and MnO2) are observed as lami-nae in travertine precipitates fromMorocco (Chafetzet al., 1998), concentrated in solution by bacterialactivity; mineral precipitation results from microen-vironmental elevation of pH and Eh conditions nearthe bacteria creating more oxidized and alkaline con-ditions suitable for Fe andMn precipitation. In addi-tion, coloured laminae observed on monumentsurfaces in Spain (Garcia-Valles et al., 1997) suggestthat trace accumulations of amorphous Fe3+ arenot part of a crystalline structure but rather associ-

ated with organic matter such as carotenoids, mela-nins, melaninoids, and humic substances. Organicstaining is also reported in calcareous crusts (Wallset al., 1975) to produce laminated horizons.

Clear differences in the elemental compositionand crystallographic form of the two laminae form-ing channel precipitates at CP-1 and CP-2 suggestdistinctly different aqueous conditions for mineralprecipitation. Chafetz et al. (1998) suggest thatalternation between the metalliferous layers and lay-ers of calcite and aragonite in Moroccan travertineresult from changes in pH within the waters result-ing from physical and chemical modification of thespring systems. Other studies show laminationsformed by seasonal changes in spring dischargerates (Drysdale, 2001), or ½PCO2

� (Kano et al.,2003). Multiple-year temperature and water chemis-try data collected from the spring outlets at ColourPeak (Pollard et al., 1998, 1999) and several years offield observations (W.H. Pollard, unpublished data)suggest that spring discharge rates and chemistryare relatively stable. It is therefore argued that sub-aerial seasonal conditions generated by the extremepolar climate of the region is responsible forchanges in aqueous chemistry and conditions formineral precipitation.

Twomajor differences in local environmental con-ditions between summer and winter months are pro-posed to affect mineral precipitation from the springwaters: (1) anoxic conditions relating to seasonalsnow cover, and (2) inputs of allochthonous debris.Furthermore, changes linked to these two conditionsoccur in spatially distinct areas to explain the lamina-tions observed both in the upstream channel precipi-tates and the rimstone pool environments.

Anoxic conditions. Evidence of moderate [HS�]combined with Eh measurements (not reported)during summer show the spring waters are anoxicat the point of emergence but become rapidly oxy-genated as they flow downslope. WDS and EDSidentification of Fe in the micritic laminae at ColourPeak provide evidence for this dissolved Fe; Fe2+

emerging from the outlets rapidly oxidizes and isbound as Fe3+ (e.g., Fe(OH)3 Æ nH2O) in the upperdischarge channels, effectively retarding or inhibit-ing calcite precipitation. Iron is not detected byICP-AES, but Fe3+ is known to rapidly oxidizeand precipitate from solution with pH > 3 in oxicenvironments (Ferris et al., 1989) and this likely oc-curred due to the delay in acidification of samples.This is in comparison to winter months, when snowcovers the upper channel areas creating a barrier

C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15 13

that reduces air–water exchange rates; Fe2+ oxida-tion is limited and calcite precipitation dominates.

Allochthonous inputs. Omelon et al. (2001) pro-pose that mineral accumulation in the lower rim-stone pools at CP-1 and CP-2 occurs duringwinter months with the precipitation of euhedralCaCO3 crystals. Observation of dissolution of thesecrystals during summer due to increased water tem-peratures is concomitant with significant inputs ofleaf litter and sediment debris to the system. Soiland organic matter effectively halt calcite precipita-tion by physical coating of crystal surfaces but actas nucleation sites for renewed calcite crystal growththe following winter. In addition, Fe3+ inputs (fromthe source waters or decaying organic matter) willalso aid the inhibition of mineral precipitation.The combined outcome of these processes producesobserved seasonal laminae alternating between theporous calcite accumulations and the horizons com-posed of allochthonous debris. Physical disruptionto mineral precipitation also occurs at GH-1, wherelayers of gypsum precipitate are interspersed withthin accumulations of leaf litter and sediments.

6. Conclusions

The springs at Expedition Fjord are uniquehydrologic systems in that they provide an opportu-nity to observe dynamic changes in chemical equi-librium as a result of atmospheric exposuredominated by cold, polar conditions. At ColourPeak, CO2 degassing from the spring waters leadsto travertine precipitation in winter months. In con-trast, gypsum precipitates from the spring waters atGypsum Hill due to lower alkalinity and higher SO4

concentrations. Mineral precipitation rates areslower than observed in other spring systems dueto the high salinities of the waters. The developmentof seasonal laminations composed of both Fe-richmicritic layers and accumulations of allochthonousleaf litter and sediment debris result from two differ-ent processes occurring throughout the summermonths. At Colour Peak, rapid oxidation of thewaters during summer retards calcite precipitationin the upper channel regions by introduction ofinhibitors such as Fe3+. In addition, inputs of allo-chthonous debris such as leaf litter and soil sedi-ment interrupts crystal growth – calcite or gypsum– at both Colour Peak and Gypsum Hill by physicalcoating of these surfaces, which act as new sub-strates for nucleation and subsequent mineral pre-cipitation the following season. Spring discharge

and associated mineral precipitates at ExpeditionFjord are a unique example of how polar conditionscan effectively control natural systems, and expandour understanding of the more complex interactionsoccurring in natural aqueous environments.

Acknowledgments

The authors acknowledge support provided by theCanadian Polar Continental Shelf Project (PCSPmanuscript 02101), the McGill University High Arc-tic Research Station, and the Natural Sciences andEngineering Research Council of Canada. Addi-tional funding for student research was provided bythe Department of Indian and Northern AffairsNorthern Scientific Training Program and McGillUniversity�s Centre for Climate and Global ChangeResearch. We thankDr. Henry Chafetz, Dr. AlfonsoMucci, and Dr. Allan Pentecost for thoughtful con-tributions to the development of this manuscript.Dr. David Parkhurst and Dr. Giles Marion areacknowledged for discussions regarding geochemicalmodeling, and Dr. S. Grasby and Dr. R. Perez arethanked for their constructive comments.

References

Amundson, R., Kelly, E., 1987. The chemistry and mineralogy ofa CO2-rich travertine depositing spring in the CaliforniaCoast Range. Geochim. Cosmochim. Acta 51, 2883–2890.

Andersen, D.T., Pollard, W.H., McKay, C., Heldmann, J., 2002.Cold springs in permafrost on Earth and Mars. J. Geophys.Res. Planets 107 (E3), 5015.

Barnes, I., 1965. Geochemistry of Birch Creek, Inyo County,California: a travertine depositing creek in an arid climate.Geochim. Cosmochim. Acta 29, 85–112.

Berner, R.A., 1981. Kinetics of weathering and diagenesis. In:Lasaga, A.C., Kirkpatrick, R.J. (Eds.), Kinetics of Geochem-ical Processes, Reviews in Mineralogy, vol. 8. MineralogicalSociety of America, pp. 111–132.

Beschel, R.E., 1963. Sulphur springs at Gypsum Hill. In:Preliminary report 1961–1962. Axel Heiberg Island ResearchReport, Miscellaneous Papers No. 2, pp. 67–193.

Bono, P., Dreybrodt, W., Ercole, S., Percopo, C., Vosbeck, K.,2001. Inorganic calcite precipitation in Tartare karstic spring(Lazio, central Italy): field measurements and theoreticalprediction on depositional rates. Environ. Geol. 41, 305–313.

Brown, R.J.E., 1972. Permafrost in the Canadian Arctic Archi-pelago. Z. Geomorphol. Suppl. Band 13, 102–130.

Bottrell, S.H., Raiswell, R., Leosson, M.A., 1996. The influenceof sulphur redox reactions and mixing on the chemistry ofshallow groundwaters: the Harrogate mineral waters. J. Geol.Soc. 153, 231–243.

Chafetz, H.S., Folk, R.L., 1984. Travertines: depositional mor-phology and the bacterially constructed constituents. J. Sed.Petrol. 54, 289–316.

14 C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15

Chafetz, H.S., Rush, P.F., Utech, N.M., 1991a. Microenviron-mental controls on mineralogy and habit of CaCO3 precip-itates: an example from an active travertine system.Sedimentology 38, 107–126.

Chafetz, H.S., Utech, N.M., Fitzmaurice, S.P., 1991b. Differencesin the d18O and d13C signatures of seasonal laminaecomprising travertine stromatolites. J. Sed. Petrol. 61, 1015–1028.

Chafetz, H.S., Akdim, B., Julia, R., Reid, A., 1998. Mn- and Fe-rich black travertine shrubs: bacterially (and nanobacterially)induced precipitates. J. Sed. Res. 68, 404–412.

Cline, J.D., 1969. Spectrophotometric determination of hydro-gen sulfide in natural waters. Limnol. Oceanogr. 14, 454–458.

Dandurand, J.L., Gout, R., Hoefs, J., Menschel, G., Schott, J.,Usdowski, E., 1982. Kinetically controlled variations of majorcomponents and carbon and oxygen isotopes in a calcite-precipitating spring. Chem. Geol. 36, 299–315.

Doran, P.T., McKay, C.P., Adams, W.P., English, M.C.,Wharton Jr., R.A., Meyer, M.A., 1996. Climate forcing andthermal feedback of residual ice covers in the high Arctic.Limnol. Oceanogr. 41, 839–848.

Dreybrodt, W., Buhmann, D., Michaelis, J., Usdowski, E., 1992.Geochemically controlled calcite precipitation by CO2 out-gassing: field measurements of precipitation rates in compar-ison to theoretical predictions. Chem. Geol. 97, 285–294.

Drysdale, R.N., 2001. Factors controlling the hydrochemistry ofLouie Creek, a travertine-depositing stream in the seasonallywet tropics of northern Australia. Mar. Freshwater Res. 52,793–804.

Emeis, K.C., Richnow, H.H., Kempe, S., 1987. Travertineformation in Plitvice National Park, Yugoslavia: chemicalversus biological control. Sedimentology 34, 595–609.

Ferris, F.G., Shotyk, W., Fyfe, W.S., 1989. Mineral formationand decomposition by microorganisms. In: Beveridge, T.J.,Doyle, R.J. (Eds.), Metal Ions and Bacteria. John Wiley &Sons, New York, pp. 413–441.

Folk, R.L., 1993. SEM imaging of bacteria and nannobacteriain carbonate sediments and rocks. Sedimentology 63, 990–999.

Folk, R.L, Chafetz, H.S., Tiezzi, P.A., 1985. Bizarre forms ofdepositional and diagenetic calcite in hot-spring travertines,central Italy. In: Schneidermann, N., Harris, P.M. (Eds.),Carbonate Cements, vol. 36. SEPM Spec. Publ., pp. 349–369.

Ford, T.D., Pedley, H.M., 1996. A review of tufa and travertinedeposits of the world. Earth Sci. Rev. 41, 117–175.

Fouke, B.W., Farmer, J.D., Des Marais, D.J., Pratt, L., Sturchio,N.C., Burns, P.C., Discipulo, M.K., 2000. Depositional faciesand aqueous-solid geochemistry of travertine-depositing hotsprings (Angel Terrace, Mammoth hot springs, YellowstoneNational Park, USA). J. Sed. Res. 70, 565–585.

Friedman, I., 1970. Some investigations of the deposition oftravertine from Hot Springs – I. The isotopic chemistry of atravertine-depositing spring. Geochim. Cosmochim. Acta 34,1303–1315.

Garcia-Valles, M., Vendrell-Saz, M., Krumbein, W.E., Urzi, C.,1997. Coloured mineral coatings on monument surfaces as aresult of biomineralization: the case of the Tarragonacathedral (Catalonia). Appl. Geochem. 12, 255–266.

Grasby, S., 2003. Naturally precipitating vaterite (l-CaCO3)spheres: unusual carbonates formed in an extreme environ-ment. Geochim. Cosmochim. Acta 67, 1659–1666.

Guo, L., Riding, R., 1992. Aragonite laminae in hot watertravertine crusts, Rapolano Terme, Italy. Sedimentology 39,1067–1079.

Guo, L., Riding, R., 1994. Origin and diagenesis of Quaternaryshrub fabrics, Rapolano Terme, central Italy. Sedimentology41, 499–520.

Hamilton, S.M., Michel, F.A., Jefferson, C.W., 1991. CO2-richground waters of the Flat River Valley, N.W.T. In: Prowse,T., Ommanney, C. (Eds.), Northern Hydrology SelectedPerspectives, Proceedings of the Northern Hydrology Sym-posium, 10–12 July 1990, Saskatoon, Saskatchewan, NHRIPaper #6, pp. 105–199.

Herman, J.S., Lorah, M.M., 1987. CO2 outgassing and calciteprecipitation in Falling Spring Creek, Virginia, USA. Chem.Geol. 62, 251–262.

Herman, J.S., Lorah,M.M, 1988. Calcite precipitation rates in thefield: measurement and prediction for a travertine-depositingstream. Geochim. Cosmochim. Acta 52, 2347–2355.

Hoen, E.W., 1964. The anhydrite diapirs of central western AxelHeiberg Island. In: Geology No. 2, Axel Heiberg IslandResearch Reports, McGill University, Montreal.

Hoffer-French, K.J., Herman, J.S., 1989. Evaluation of hydro-logical and biological influences on CO2 fluxes from a karststream. J. Hydrol. 108, 189–212.

Jacobson, R.L., Usdowski, E., 1975. Geochemical controls on acalcite precipitating spring.Contrib.Mineral. Petrol. 51, 65–74.

Jones, B., Renaut, R.W., 1996. Morphology and growth ofaragonite crystals in hot-spring travertines at Lake Bogoria,Kenya Rift Valley. Sedimentology 43, 323–340.

Kano, A., Matsuoka, J., Kojo, T., Fujii, H., 2003. Origin ofannual laminations in tufa deposits, southwest Japan. Palae-ogeogr. Palaeocl. Palaeoecol. 191, 243–262.

Katz, J.L., Reick, M.R., Herzog, R.E., Parsiegla, K.I., 1993.Calcite growth inhibition by iron. Langmuir 9, 1423–1430.

Langmuir, D., 1997. Aqueous Environmental Chemistry. Pre-ntice Hall, New Jersey.

Lauritzen, S.E., Bottrell, S., 1994. Microbiological activity inthermoglacial karst springs, South Spitzbergen. Geomicro-biol. J. 12, 161–173.

Liu, Z., Svensson, U., Dreybrodt, W., Daoxian, Y., Buhmann,D., 1995. Hydrodynamic control of inorganic calcite precip-itation in Huanglong Ravine, China: field measurements andtheoretical prediction of deposition rates. Geochim. Cosmo-chim. Acta 59, 3087–3097.

Liu, Z., Zhang, M., Li, Q., You, S., 2003. Hydrochemical andisotope characteristics of spring water and travertine in theBaishuitai area (SW China) and their meaning for paleoen-vironmental reconstruction. Environ. Geol. 44, 698–704.

Lorah, M.M., Herman, J.S., 1988. The chemical evolution of atravertine-depositing stream: geochemical processes and masstransfer reactions. Water Resour. Res. 24, 1541–1552.

Losh, S., Walter, L., Meulbroek, P., Martini, A., Cathles, L.,Whelan, J., 2002. Reservoir fluids and their migration into theSouth Eugene Island Block 330 reservoirs, offshore Louisiana.AAPG Bull. 86, 1463–1488.

Love, K.M., Chafetz, H.S., 1988. Diagenesis of laminatedtravertine crusts, Arbuckle Mountains, Oklahoma. J. Sed.Petrol. 58, 441–445.

Matsuoka, J., Kano, A., Oba, T., Watanabe, T., Sakai, S., Seto,K., 2001. Seasonal variation of stable isotopic compositionsrecorded in a laminated tufa, SW Japan. Earth Planet. Sci.Lett. 192, 31–44.

C.R. Omelon et al. / Applied Geochemistry 21 (2006) 1–15 15

Meyer, H.J., 1984. The influence of impurities on the growth rateof calcite. J. Cryst. Growth 66, 639–646.

Mitterer, R.M., Cunningham Jr., R., 1985. The interaction ofnatural organic matter with grain surfaces: implications forcalcium carbonate precipitation. In: Schneidermann, N.,Harris, P.M. (Eds.), Carbonate Cements, vol. 36. SEPMSpec. Publ., pp. 17–31.

Morse, J.W., 1983. The kinetics of calcium carbonate dissolutionand precipitation. In: Reeder, R.J. (Ed.), Carbonates: Min-eralogy and Chemistry, Reviews in Mineralogy, vol. 2.Mineralogical Society of America, pp. 227–264.

Morse, J.W., Mucci, A., Walter, L.M., 1979. Magnesiuminteraction with the surface of calcite in seawater. Science205, 904–905.

Nassichuk, W.W., Davies, G.R., 1980. Stratigraphy and sedi-mentation of the Otto Fiord Formation – a major Mississip-pian-Pennsylvanian evaporite of subaqueous origin in theCanadian Arctic Archipelago. Geol. Surv. Can. Bull., 286.

Omelon, C.R., Pollard, W.H., Marion, G.M., 2001. Seasonalformation of ikaite (CaCO3 Æ 6H2O) in saline spring dischargeat Expedition Fiord, Canadian High Arctic: assessing condi-tional constraints for natural crystal growth. Geochim.Cosmochim. Acta 65, 1429–1437.

Pentecost, A., 1992. Carbonate chemistry of surface waters in atemperate karst region: the southern Yorkshire Dales, UK. J.Hydrol. 139, 211–232.

Pentecost, A., 1995. Geochemistry of carbon dioxide in sixtravertine-depositing waters of Italy. J. Hydrol. 167, 263–278.

Pentecost, A., 1996. Moss growth and travertine deposition: thesignificance of photosynthesis, evaporation and degassing ofcarbon dioxide. J. Bryol. 19, 229–234.

Pentecost, A., Bayari, S., Yesertener, C., 1997. Phototrophicmicroorganisms of the Pamukkale Travertine, Turkey: theirdistribution and influence on travertine deposition. Geomi-crobiol. J. 14, 268–283.

Pentecost, A., Jones, B., Renaut, R.W., 2003. What is a hotspring? Can. J. Earth Sci. 40, 1443–1446.

Plummer, L.N., Parkhurst, D.L., Fleming, G.W., Dunkle, S.A.,1988. A computer program incorporating Pitzer�s equationsfor the calculation of geochemical reactions in brines. USGeological Survey Water-Resources Investigations Report88-4153.

Pollard, W.H., Omelon, C., Andersen, D., McKay, C., 1998.Geomorphic and hydrologic characteristics of perennialsprings on Axel Heiberg Island, Canadian High Arctic. In:Proceedings of the 7th International Conference on Perma-frost, Canadian National Committee for the InternationalPermafrost Association (IPA), Yellowknife, N.W.T., June23–27, 909-914.

Pollard, W.H., Omelon, C., Andersen, D., McKay, C., 1999.Perennial spring occurrence in the Expedition Fiord area of

western Axel Heiberg Island, Canadian High Arctic. Can. J.Earth Sci. 36, 105–120.

Reddy, M.M., 1977. Crystallization of calcium carbonate in thepresence of trace concentrations of phosphorus-containingions. I. Inhibition of phosphate and glycerophosphate ions atpH 8.8 and 25 �C. J. Cryst. Growth 41, 287–295.

Reddy, M.M., 1983. Characterization of calcite dissolution andprecipitation using an improved experimental technique. Sci.Geol. Mem. 71, 109–117.

Reddy, M.M., Nancollas, G.H., 1976. The crystallization ofcalcium carbonate. IV. The effect of magnesium, strontium,and sulfate ions. J. Cryst. Growth 35, 33–38.

Riding, R., 2000. Microbial carbonates: the geologic record ofcalcified bacterial–algal mats and biofilms. Sedimentology 47,179–214.

Taylor, A.E., Judge, A., 1976. Canadian Geothermal DataCollection. Geothermal Series Number 6, Earth PhysicsBranch, EMR, Ottawa.

Thorsteinsson, R., 1974. Carboniferous and Permian stratigraphyof Axel Heiberg Island and western Ellesmere Island, Cana-dian Arctic Archipelago. Geol. Surv. Can. Bull. 224, 1–115.

Tijani, M.N., 2004. Evolution of saline waters and brines in theBenue-Trough, Nigeria. Appl. Geochem. 19, 1355–1366.

Usdowski, E., Hoefs, J., 1990. Kinetic 13C/12C and 18O/16Oeffects upon dissolution and outgassing of CO2 in the systemCO2–H2O. Chem. Geol. 80, 109–118.

Usdowski, E., Hoefs, J., Menschel, G., 1979. Relationshipbetween 13C and 18O fractionation and changes in majorelement composition in a recent calcite-depositing spring – amodel of chemical variations with inorganic CaCO3 precip-itation. Earth Planet. Sci. Lett. 42, 267–276.

van Everdingen, R.O., Shakur, M.A., Krouse, H.R., 1982.Isotope geochemistry of dissolved, precipitates, airborne,and fallout sulfur species associated with springs near PaigeMountain, Norman Range, N.W.T. Can. J. Earth Sci. 19,1385–1407.

Walls, R.A., Harris, W.B., Nunan, W.E., 1975. Calcareous crust(caliche) and early subaerial exposure of Carboniferous car-bonates, northeasternKentucky. Sedimentology 22, 417–440.

Weijermars, R., Mulder-Blanken, C.W., Wiegersm, J., 1986.Growth rate observation from the moss-built Checa traver-tine terrace, central Spain. Geol. Mag. 123, 279–286.

Zuber, A., Weise, S.M., Osenbruck, K., Matenko, T., 1997.Origin and age of saline waters in Busko Spa (southernPoland) determined by isotope, noble gas and hydrochemicalmethods: evidence of interglacial and pre-Quaternary warmclimate recharges. Appl. Geochem. 12, 643–660.

Zuddas, P., Mucci, A., 1994. Kinetics of calcite precipitationfrom seawater. I. A classical chemical kinetics description forstrong electrolyte solutions. Geochim. Cosmochim. Acta 58,4353–4362.