Quick viewing(Text Mode)

The Ordovician: a Window Toward Understanding Abundance

The Ordovician: a Window Toward Understanding Abundance

THE : A WINDOW TOWARD UNDERSTANDING ABUNDANCE

AND MIGRATION PATTERNS OF BIOGENIC AND IMPLICATIONS FOR

PALEOCLIMATE

A thesis presented to

the faculty of

the College of Arts and Sciences of Ohio University

In partial fulfillment

of the requirements for the degree

Master of Science

Iulia Tomescu

August 2004 © 2004

Iulia Tomescu

All Rights Reserved This thesis entitled

THE ORDOVICIAN: A WINDOW TOWARD UNDERSTANDING ABUNDANCE

AND MIGRATION PATTERNS OF BIOGENIC CHERT AND IMPLICATIONS FOR

PALEOCLIMATE

BY

IULIA TOMESCU

has been approved for

the Department of Geological Sciences

and the College of Arts and Sciences by

David L. Kidder

Associate Professor of Geological Sciences

Leslie A. Flemming

Dean, College of Arts and Sciences TOMESCU, IULIA. M.S. August 2004. Geological Sciences

The Ordovician: a window toward understanding abundance and migration patterns of

biogenic chert and implications for paleoclimate (228 pp.)

Director of Thesis: David L. Kidder

The comprehensive global survey of Ordovician chert deposits undertaken in this study reveals that most of them are biogenic, and documents significant paleoceanographic relationships. Peritidal/lagoonal and shelf hosted mainly by facies are nodular, and formed predominantly by lithistid .

Slope/basinal cherts are bedded, hosted by siliciclastic , and bear radiolarians.

Lower Ordovician cherts are abundant in all settings. The Middle Ordovician shows a decrease in cherts characterized by a strong decline in silica burial within peritidal/lagoonal and shelf environments. The distribution of silica secreting biotas suggests an Arenigian- Llanvirnian onshore-offshore migration whereby siliceous migrate from peritidal/lagoonal to shelf-basinal environments, and radiolarians move to basinal settings. The Caradocian marks the highest peak in the abundance of

Ordovician cherts, in both shelf and slope/basinal environments, that appears to correlate with a warming interval. A decline of cherts takes place in the Ashgillian, when glaciation was in effect.

Approved:

David L. Kidder

Professor of Geological Sciences Acknowledgments

My sincere thanks go to Dr. David Kidder, my mentor. He introduced me to the silica system and all its biotic and abiotic components. During our long discussions he

kindly guided me through the maze of the Ordovician Earth system and Paleozoic silica-

secreting organisms. I am very grateful for having had the chance to work with him.

I would also like to thank Drs. Thomas Worsley and Gregory Nadon. Dr. Worsley

opened the world of Earth system evolution to me. His broad-picture integrative models

were a constant source of inspiration, and his feed-back greatly encouraged me in my

scholarly pursuits. Dr. Nadon was always available for discussions and interactions with

him sharpened my critical thinking. 6

Table of Contents

Page Abstract...... 4

Acknowledgments...... 5

List of Tables ...... 8

List of Figures...... 9

Chapter 1. Significance of the study...... 11

Chapter 2. Hypotheses of the study ...... 13

Chapter 3. Assumptions of the study ...... 17

Chapter 4. Previous work...... 19

Chapter 5. Bedded versus nodular cherts...... 27 5.1 What is chert? ...... 27 5.2 Chert formation...... 30 5.3 Chert deposit types: characteristics and genesis theories ...... 32 5.4 Time and temperature of certification and depth of formation...... 40

Chapter 6. Methodology ...... 42 6.1 Data acquisition ...... 42 6.2 Data management...... 43 6.3 Data structure...... 43 6.4 Data manipulation...... 45 6.5 Pitfalls/Problems of a study compiling data from published reports...... 54 6.6 The usefulness of a study compiling data from published reports...... 57

Chapter 7. Results ...... 58 7.1 Epoch - level temporal trend in the chert abundance...... 58 7.2 Epoch - level temporal trend in chert types ...... 60 7.3 The spatio-temporal distribution of chert occurrences ...... 63 7.4 Chert type - depositional environment relationship over time...... 77 7.5 The temporal and spatial distribution of the Ordovician silica secreting organisms present in chert deposits ...... 78 7.6 Testing the significance of associations between the type of siliceous skeletal remains, chert deposit styles, and depositional environments...... 82 7.7 Evaluation of the efficiency of the method...... 83

7

Table of Contents: continued.

Page Chapter 8. The Lower Paleozoic silica secreting organisms: radiolarians and siliceous sponges...... 91 8.1 Radiolarians ...... 91 8.2 Siliceous sponges...... 101

Chapter 9. The origin of the Ordovician cherts - organic versus inorganic genesis...... 111 9.1 Previous theories on chert origin ...... 111 9.2 Evidence of the present study ...... 116 9.3 Silica burial: a balance between silica dissolution and preservation rates.....122

Chapter 10. Discussion: the Ordovician silica secreting biotas - living style, distribution and onshore-offshore evolution patterns ...... 127 10.1 Lower Ordovician ...... 127 10.2 Middle Ordovician...... 136 10.3 Upper Ordovician...... 141

Chapter 11. Interpreting the chert abundance patterns: why chert peaks or lows?...... 143 11.1 Lower Ordovician: a greenhouse interval...... 143 11.2 Middle Ordovician: a greenhouse interval...... 155 11.3 Upper Ordovician: the Caradocian - a warm - cool transitional, ice free mode (?) ...... 160 11.4 Upper Ordovician: the Ashgillian - an icehouse period...... 173

Chapter 12. Conclusions ...... 182

References...... 187

Appendix ...... 220

8

List of Tables

Table ...... Page

7.1 Temporal distribution of the chert deposits at epoch and age/stage level...... 59

7.2 Temporal distribution of the chert deposit types at epoch (A) and stage (B) level. ....61

7.3 Temporal (A epoch level; B stage level) and spatial (Onshore – Offshore) distribution of the chert deposits...... 64

7.4 Distribution of the chert deposit types by epoch and depositional environment...... 72

7.5 Temporal trends in the paleolatitudinal distribution of Ordovician silica burial...... 75

7.6 Temporal and geographical distribution of chert deposits...... 76

7.7 Temporal distribution of Ordovician silicified oolites...... 76

7.8 Epoch level distribution of chert deposits with siliceous skeletal remains...... 79

7.9 Relationships between chert deposit types and types of siliceous skeletal remains for the Lower, Middle, and Upper Ordovician...... 80

9

List of Figures

Figure ...... Page

4.1 Ordovician paleogeography, chert distribution, and oceanic circulation...... 21

4.2 Phanerozoic distribution of bedded cherts compiled at the system level...... 23

4.3 Phanerozoic distribution of bedded cherts compiled at the system level...... 23

4.4 Phanerozoic distribution of bedded cherts compiled at the series (epoch) level...... 24

4.5 Sedimentary silica through time...... 25

4.6 (A) Temporal distribution of silicified oolites and bedded cherts at the series level.(B) Onshore-offshore migration patterns of three faunas...... 26

6.1 FileMaker database entry sample...... 44

6.2 Ordovician correlation chart used in this study...... 52

7.1 Ordovician distribution of chert deposits...... 59

7.2A Temporal distribution of chert deposit types at the epoch level...... 62

7.2B Temporal variation of the relative abundance of chert deposit types at the epoch level...... 62

7.3 Ordovician spatio-temporal distribution of chert occurrences...... 65

7.4 Paleogeographic distribution of Lower Ordovician cherts...... 66

7.5 Paleogeographic distribution of Middle Ordovician cherts...... 68

7.6 Paleogeographic distribution of Upper Ordovician cherts...... 70

7.7 Relationships between chert types and depositional environments over time...... 73

7.8 Epoch-level variation of the relative abundance of chert formations that bear siliceous skeletal remains...... 79

10

List of Figures: continued.

7.9 Chi-square of the relationship between the type of siliceous organisms...... 84

7.10 Chi-square test of the relationship between the siliceous organisms...... 85

7.11 Chi-square test of the relationship between the depositional environment and the chert deposit style...... 87

10.1 Ordovician dominant locus of silica burial and onshore-offshore migration patterns of silica secreting organisms and associated facies...... 128

11.1A The Ordovician biotic system...... 144

11.1B The Ordovician abiotic system...... 146

11.2 Lower Ordovician - Chert abundance model...... 151

11.3 Middle Ordovician - Chert abundance model...... 156

11.4 Temperature vs. calcium carbonate shell δ18O (PDB) for various values of seawater δ18O (SMOW) in an ice-cap free world...... 164

11.5 Upper Ordovician: Caradocian - Chert abundance model...... 168

11.6 Upper Ordovician: Ashgillian - Chert abundance model...... 177 11

Chapter 1

THE SIGNIFICANCE OF THE STUDY

The present study represents a unique comprehensive systematic survey of the

worldwide Ordovician early diagenetic in situ chert deposits, addressing not only bedded

cherts, but also less conspicuous cherty facies such as nodular cherts and associated

silicified oolites. The distinction of the chert deposit types (nodular cherts vs. bedded

cherts vs. silicified oolites), the depositional environment in which cherts formed

(peritidal/lagoonal vs. shelf vs. slope/basin), and a systematic analysis of the content in

siliceous skeletal remains (radiolarians and siliceous sponges) preserved in the compiled

chert deposits, were among the main targets in surveying the data. The compilation was

realized with a high temporal resolution at the epoch level, and where possible, even at

the stage level (e.g., the Caradocian).

This study is the first qualitative and quantitative test of two common

assumptions: 1) the biogenic origin hypothesis for most of the Ordovician cherts and a

biologically controlled marine and burial and, 2) the relationship between the

chert deposit types (nodular vs. bedded chert) and the hosting facies (carbonate vs. siliciclastic environments). This research also evaluates: 1) the relationship between the chert deposit types (nodular vs. bedded) and the depositional environment in which they formed (peritidal/lagoonal/shelf vs. slope/basinal); and 2) the relationship between the chert deposit types (nodular vs. bedded) and the content in biogenic siliceous remains

12

(radiolarians vs. siliceous sponges). The results demonstrate that the Ordovician onshore- offshore organism migration, a mechanism suggested to date only for the benthic invertebrate macrofauna, included planktonic radiolarians and siliceous sponges as well.

The study goals were: 1) to identify and interpret Ordovician temporal and spatial patterns in the locus of marine silica accumulation in fluctuations of the marine biogenic chert abundance, in possible migrations of the silica secreting organisms, 2) to correlate these patterns with paleoclimate, paleoceanography, paleogeography, level inputs, and the behavior of silica secreting organisms, and 3) to translate these interconnections in terms of causal relationship. Through this, the study represents a multi-approach analysis of the Ordovician Earth system, whose main goal is to explore the connections between the biotic and abiotic worlds.

13

Chapter 2

THE HYPOTHESES OF THE STUDY

Based on previous work (discussed in Chapter 4) the following hypotheses have

been proposed to be tested:

I. Did the Early Paleozoic radiolarians and siliceous sponges behave similarly to

regarding their response to abrupt changes in oceanic circulation and climate

parameters?

Assessing the worldwide abundance and distribution of Early Paleozoic cherts in

relation to paleoclimatic events will help clarify whether radiolarians and siliceous

sponges were as responsive to oceanographic changes as are diatoms. This behavior was

tested for the Ordovician, a characterized by a succession of

antagonistic climates: Lower and Middle Ordovician warm greenhouse periods, as

opposed to the late Upper Ordovician, an icehouse interval. A chert peak is expected to be associated with an environment driven by the late Ordovician glaciation. In

contrast, a gap in chert abundance in the Early may reflect sluggish

circulation and possible climate warming similar to the Middle and Early

Triassic, both of which are marked by lows in chert abundance.

The accumulation and formation of cherts depends on different factors –

upwelling is very important in offshore environments, while in nearshore facies other

14 factors are important, such as the amount of dissolved silica supplied to seawater by riverine input and by the siliceous organisms that inhabit these areas (e.g., siliceous sponges). Consequently, during glacial periods bedded chert accumulation should be common in offshore settings, driven by an enhanced upwelling. On the contrary, in nearshore facies the amount of siliceous deposits could decrease as a consequence of changes in the ecological distribution of siliceous organisms due to sea level fluctuations.

Changes in chert distribution could be driven by a drop in sea level appreciated to have amounted to at least 50m for the Late Ordovician glaciation. On the other hand, a drop in sea level would increase the volume of terrestrial available to , and possibly the dissolved silica input into the basinal facies, while the shelf would witness decreased deposition being more or less emergent. The amount of dissolved silica that enters the via rivers (the exogenous silica input) is directly connected to the continental weathering rates, whose intensity is a result of the interplay between the rock volume exposed to weathering, atmospheric conditions (temperature, precipitations), and the average paleolatitudinal distribution of the world landmasses. If the interaction between the three parameters drives a higher weathering rate, the exogenous, dissolved silica input in coastal seawaters will be higher resulting in more chert accumulation in areas of low detrital influx.

15

II. The timing of the Late Ordovician glaciation - Did a brief glaciation result in a

short cherty interval? Did cherts respond quickly enough to reflect the duration of the

glaciation?

The Late Ordovician is characterized by a glaciation the timing of which is open

to controversy. Brenchley et al. (1994) suggest that the late Ordovician glaciation may have lasted less than 1 m.y. during an otherwise long greenhouse interval. Conversely,

Pope and Steffen (2001) suggest that extensive chert in the Montoya Group represents a glacial-related upwelling system that persisted for up to 15 m.y.

If the Paleozoic siliceous organisms can be demonstrated to be sensitive to climatic and oceanographic variations, then we need to know the degree of this sensitivity. A short-lived chert pulse, coeval with the glaciation, would support the short

glaciation hypothesis, and would argue that Ordovician biogenic cherts are very sensitive

to abrupt changes in ocean circulation. A longer interval of chert deposition would be

consistent with a glaciation of longer duration or in the case of a short glaciation, might

mean that siliceous sponges and radiolarian cherts were not as responsive to

oceanographic changes as are cherts. If glaciation was short, an alternative

explanation for abundant, pre-glacial chert in the Late Ordovician is required.

III. The Early Paleozoic dominant locus of silica deposition migrated from shallow

water to deeper water environments within the Ordovician period.

Analysis of the Ordovician spatial and temporal distribution of chert sedimentary

facies, such as bedded, nodular, and silicified oolite deposits, can be used to: 1) estimate

16 the dominant locus of silica deposition, onshore vs. offshore; 2) infer onshore-offshore migration patterns of silica secreting organisms perceived through changes in the depositional environment of biogenic siliceous facies; 3) analyze the timing, the rate and the nature of this retreat of chert to deeper water settings; and 4) identify and interpret marine chert abundance fluctuations and patterns.

17

Chapter 3

THE ASSUMPTIONS OF THE STUDY

The theory according to which seawater was undersaturated with respect to amorphous silica not only during the and , but also during the

Paleozoic, is widely accepted today (Siever, 1957; Grunau 1965; Hein and Parrish, 1987;

Maliva et al., 1989). Paleozoic marine burial of silica has been controlled for most of the time by the Paleozoic silica secreting organisms represented by radiolarians and siliceous sponges. In a marine silica cycle in which silica burial is dominated by the biogenic component, the dissolution of the siliceous , which takes place during the living stage of the organisms (Spencer, 1983; De la Rocha et al., 1997) as well as within the while dead organisms are sinking and continues during burial (DeMaster,

1981), represents the main mechanism for remobilizing and recycling dissolved silica in the sea water (Tréguer et al., 1995). The silica secreting biotas, due to their biochemical efficiency of silica uptake, are responsible for keeping the seawater undersaturated with respect to amorphous silica, and thus preventing the precipitation of silica and the formation of extensive inorganic cherts.

The Ordovician chert deposits are considered in this study to be mainly biogenic in origin, induced either organically (based on biological silica recycling), or inorganically (based on inorganic silica and other nutrient inputs that drive biological blooms of the silica secreting organisms). Deposits demonstrated to be inorganically

18 generated (e.g., chert units associated with tephra layers or bentonites, and possibly the bedded cherts formed in peritidal/lagoonal environments characterized by restrictive circulation) are exceptions. Biogenic chert deposits are considered those for which the bulk silica input originates in the skeletons of silica secreting biotas, representing a net output of Si from the biochemical cycle.

Taking into account that skeletons of the siliceous organisms are the main supplier of reactive silica to sediments, the abundance of nodular and bedded cherts would reflect environments rich in such organisms, in either deep or shallow waters. In other words, “in situ” early diagenetic chert deposits are assumed to reflect the presence of silica-secreting organisms in their proximal environment, and temporal changes in the depositional environment of silica burial perceived through the rock record reflect biological variations of the silica-secreting organisms.

19

Chapter 4

PREVIOUS WORK

Previous chert compilation-based studies surveyed worldwide records of bedded chert deposits only, with the aim to identify and interpret Phanerozoic chert abundance patterns (Grunau, 1965; Dietz and Holden, 1966; Ramsay, 1973; Ronov (1982); Hein and

Parrish, 1987; Kidder and Erwin, 2001). Grunau’s (1965) work led to the suggestion that past chert abundance (such as ) may have been related to ophiolite genesis.

This theory was based on his finding that ophiolite thickness appears to correlate with the thickness of approximately coeval chert deposits. This direct relationship between ophiolite and chert formation, combined with the “eugeosynclinal” position of many of these radiolarites (Grunau, 1965) implied a volcanic source of silica. Racki and Cordey

(2000) suggest that volcanic sources of silica may have been more significant in the

Paleozoic than in younger rocks. Given the relative lack of orogenic activity in the

Cambrian and early to middle Ordovician, it is reasonable to suspect a greater relative role for mid ocean ridge silica sources. weathering associated with Late

Ordovician orogenic activity (e.g., Kump et al., 1999) probably increased the terrestrial input of silica. The bedded chert compilations of Ramsay (1973) and Ronov (1982) attempted to compare the volume of chert abundance through time along with volumes for other sedimentary rocks. Hein and Parrish (1987) expanded the work of Grunau

(1965) and Ramsay (1973), and constructed a paleo-upwelling model of bedded chert 20 distribution (fig. 4.1). All these chert data ware assembled into period-level groupings.

The high sensitivity of radiolarians to water temperature and chemistry has rendered radiolarians an excellent paleoceanographic and paleoclimatologic marker, mainly for the

Cenozoic era. Distribution patterns and abundance of distinct modern radiolarian assemblages are tightly linked to the amount of dissolved in the seawater and to water depth (Kruglikova, 1993; Abelmann and Gowing, 1997). Based on an actualistic scenario, the abundance of radiolarian-bearing strata has been used as an indicator of paleo-upwelling and high areas for the Phanerozoic (Hein and Parrish,

1987). Also, radiolarian-bearing chert deposits preserved since the Ordovician have been used successfully as an indicator of deep-water settings (distal basinal or oceanic environments associated with orogenic belts), but have been also associated with cratonic seas of moderate depths as well (Aubouin, 1965; Ormiston, 1993). Vishnevskaya (1997) has shown that evolutionary changes undergone by radiolarians throughout the

Phanerozoic could be linked to sea level changes, paleoceanographic regime, and plate tectonics events. Numerous Lower Paleozoic radiolarian assemblages are incorporated into the accretionary prisms associated with ancient subduction zones (Aitchinson, 1998).

Because numerous Paleozoic radiolarian-bearing chert deposits are associated with ophiolitic complexes, many authors emphasize the use of radiolarians as a biostratigraphic tool in dating lower Paleozoic orogenic belts devoid of other biostratigraphic markers and consequently lacking a good age control, (Grunau, 1965;

Aitchinson et al., 1998; Blome et al., 1995b; Nazarov and Ormiston 1993). 21

Figure 4.1: Ordovician paleogeography, chert distribution (after Hein and Parrish, 1987), and oceanic circulation (after Berry et al., 1995).

22

Kidder and Erwin (2001) analyzed the Phanerozoic distribution of bedded chert in order

to determine the effects of and climate changes on siliceous facies. They

refined the Hein and Parrish (1987) data to the epoch level and distinguished basinal from

shelf cherts based on facies associations. The raw abundance data at system-level shows a

marked decrease in both basinal and shelf chert during the Silurian period but the

Ordovician period displays no peak or decrease in the number of chert occurrences (fig.

4.2.). On the other hand, the chert data normalized to outcrop area/time (after Raup,

1976) suggests a small Ordovician chert peak (fig. 4.3.). The distribution of chert data compiled at the epoch level produces an Upper Ordovician chert peak (fig. 4.4.) that might be associated with the Late Ordovician glaciation through enhanced oceanic circulation that would favor stronger upwelling and implicitly higher organic productivity and chert formation rate. The difficulty with the Kidder and Erwin (2001) compilation is that the number of bedded cherts per epoch is commonly less than ten. This low sample size provides only weak support for high versus low oceanic productivity. Those results are sufficient however, to stimulate testable hypotheses.

Maliva at al. (1989), Kidder and Erwin (2001), and Kidder and Mumma (2003) analyzed the dominant locus of burial and its migration patterns throughout the Phanerzoic. The work of Maliva at al. (1989), based on the depositional distribution of chert facies, showed that the -Ordovician interval represents a transitional period in the dominant style of silica burial from peritidal-dominated environments of early diagenetic chert facies, to a pattern of bedded chert-dominated basinal and nodular chert-dominated shelf/slope environments, that they suggest to have 23

Figure 4.2: Phanerozoic distribution of bedded cherts compiled at the system level (after Kidder and Erwin, 2001). Raw abundance data. Arrows highlight Ordovician chert.

Figure 4.3: Phanerozoic distribution of bedded cherts compiled at the system level (after Kidder and Erwin, 2001). Raw chert data presented in figure 4.2 is normalized to outcrop area. Arrows highlight Ordovician chert. 24

Figure 4.4: Phanerozoic distribution of bedded cherts compiled at the series (epoch) level (after Kidder and Erwin, 2001).

25

persisted from the Silurian through the Cretaceous (fig. 4.5). During the Cenozoic the

main locus of silica burial is confined to bedded chert-dominated deep-sea settings

(Maliva et al., 1989).

The worldwide compilation of the Phanerozoic in situ silica-replaced oolites silicified during early (Kidder and Mumma, 2003), and the depositional environment analysis of shelf bedded cherts vs. basinal bedded cherts (Kidder and Erwin,

2001), suggest that the Early Paleozoic migration of cherty facies from shallow-water to deeper environments may have occurred within the Ordovician (fig. 4.6). Again, however, low sample size restricts this interpretation to the status of a testable hypothesis.

Figure 4.5: Sedimentary silica through time (after Maliva et al., 1989).

26

Figure 4.6: (A) Temporal distribution of silicified oolites and bedded cherts at the series level. Numbers at right are the percentages of shelf chert occurrences from total chert occurrences in each series. (B) Onshore-offshore migration patterns of the three faunas after Cowen’s (2000) reinterpretation of Sepkoski and Miller’s (1985) data (after Kidder and Mumma, 2003).

27

Chapter 5

BEDDED CHERTS versus NODULAR CHERTS

5.1. What is chert?

5.1.1. Definition and terminology

After siliceous organisms die, their opaline skeletons either accumulate on the sea

bottom (in the case of the planktonic siliceous fauna), or dissolve and reprecipitate (in the

case of the benthic siliceous organisms) and are subsequently transformed into chert

through complex diagenetic processes. Chert refers to any siliceous

whose origin can be inorganic (volcanic and/or hydrothermal) and organic (biochemical

or biogenic). “Siliceous sedimentary rocks are fine grained, dense, very hard rocks

composed mainly of the SiO2 such as , chalcedony, and opal, with minor

impurities such as siliciclastic grains and diagenetic minerals” (Boggs, 1987, p.239). In a

strict sense, the term chert designates a dense microcrystalline or cryptocrystalline

sedimentary rock, consisting of quartz crystals less than about 30 µm in diameter

(Calvert, 1974). Other names of chert are: , designating a dark gray or black variey of chert; jasper, a red variety of chert containing disseminated hematite; novaculite, defined as a very dense, fine-grained, even-textured chert that occurs mainly in mid-

Paleozoic rocks of the , Oklahoma, and region of the south-central United

States; porcellanite, representing a fine-grained impure siliceous rock characterized by a texture and fracture close to unglazed porcelain and composed mainly of cristobalite and

28

subordinately clay minerals; silex, representing early diagenetic siliceous nodules in

of biochemical origin, with a large variety of colors, fine grained and very

hard, composed mainly of chalcedony and cryptocrystalline quartz.

5.1.2. Mineralogy, texture and chemical composition

The main minerals that constitute the chert deposits are: amorphous silica (opal-

A), cristobalite (opal-CT and opal-C), tridymite, chalcedonic quartz, microquartz and

various combinations of these phases. Cristobalite and tridymite are polymorph varieties of quartz. According to Lancelot (1973b), the mineralogy of chert is influenced by the lithology of the host rock. His hypothesis claims that chert in clay-rich sediments (zeolitic clays, clayey radiolarian oozes, marls and marly limestones) is porcelanite (rock composed mainly of disordered cristobalite – opal-CT), whereas in calcareous oozes

chert mineralogy is exclusively represented by quartz. Based on clay mineralogy of the

bed, Lancelot suggests that chert is more prone to form within a silicic environment

relatively rich in K-felspars, and palygorskite (a chain-lattice clay with good

bleaching and adsorbent properties).

Opal-A or amorphous silica, with up to 10% water, represents the constituent

material of the in siliceous organisms. Von Rad and Rösch (1974) define opal-A

as highly disordered, nearly amorphous natural hydrous silica. It is metastable and

therefore it becomes less frequent back through time, being absent from Paleozoic cherts

(Tucker, 2001).

Opal-CT represents an intermediate phase of the opaline silica transformation

into chert (quartz). The 3-layered mineralogical structure is represented by disordered

29

stacking of a low temperature cristobalite layer and two low-temperature trydimite layers

(Wise and Weaver, 1974; Boggs, 1987). Opal-CT may occur as microcrystalline aggregates 5-15 µm in diameter called lepispheres, in turn composed of fine blade-shaped

crystals 300-500 Å thick (Calvert 1974). Cristobalite represents an important constituent

of many deep-sea cherts. In the chert literature, low temperature cristobalite can be found

under various names such as: cristobalite, α-cristobalite, disordered cristobalite, lussatite,

opal-CT, opal-cristobalite, unidimensionally disordered cristobalite (Wise and Weaver,

1974). It is considered to be an early diagenetic product of low temperatures.

Quartz in chert is a late diagenetic product and based on grain size is divided into

megaquartz (> 20 µm) and microquartz (< 20 µm). Microquartz can be of two types:

microcrystalline (1-4 µm), and chalcedonic quartz with a fibrous habit (Carson, 1991).

According to von Rad and Rösch (1974), a quartz-chert lacks and porosity.

Chalcedonic quartz is fibrous quartz of three types, depending on the optical

orientation of the fibers: two length-slow types (quartzine and lutecite), and one length-

fast type of chalcedony (chalcedonite). The latter is considered a late diagenetic product

(Heath, 1974).

From a chemical point of view, in addition to silica (SiO2), cherts contain minor

amounts of Al, Fe, Mn, Ca, Na, K, Mg, plus some impurities such as authigenic hematite

and pyrite, siliciclastic minerals, or pyroclastic particles. Aluminum represents commonly

the second most abundant element in cherts, followed by Fe, Mg, or K, Ca, and K. The

amount of silica in chert ranges from 99% for a pure chert (such as the Arkansas

Novaculite), to less than 65% in some chert nodules (Boggs, 1987).

30

5.2. Chert formation

Boggs (1987, p.92) maintained that chert formation is a dual process: 1) a

sedimentary process reflecting the accumulation and concentration of biogenic opaline

tests by burial and, 2) a diagenetic process reflecting crystallization and recrystallization of the amorphous silica after burial.

The amount of dissolved silica in the pore water of marine sediments is higher than that of adjacent bottom waters (Tréguer et al., 1995). Within the , silica is a mobile component: a part is recycled into the overlying water by dissolution, another part

is involved in the formation of aluminosilicate mineral phases, and the rest crystallizes

and is preserved as chert. From the total biogenic silica production, 97.5% is recycled and

only 2.5% is preserved in the sediment as chert (Tréguer et al., 1995). According to

Heath (1974), 4% of the dissolved silica is buried, out of which only 2% will be left after

the post-depositional dissolution, and preserved as chert. Within the sediment high

dissolved silica concentrations correlate with biogenic silica-rich sediments. Dissolved silica increases rapidly in concentration in the uppermost layers of the sediment, and reaches a constant value some tens of centimeters below the surface of the sediment

(Calvert, 1983).

Von Rad and Rösch (1974) contend that the first steps in chertification are the remobilization of dissolved biogenous opal-A and precipitation of lussatite blades (low temperature cristobalite), with formation of microspherules (an early diagenetic process), followed by the conversion of the lussatite into chalcedony and/or micro- (>1 µm) to cryptocrystalline (< 1 µm) quartz (a late diagenetic processes).

31

Rapid siliceous test accumulation results in a high opal concentration per surface

unit area. Burial and continuous dissolution of the opaline skeletal material increasingly enriches the pore waters in silica favoring a slow precipitation of chert. The most widely accepted diagenetic progression is the maturation theory (Kastner et al., 1977) in which opal-A passes into opal-CT (disordered cristobalite) that in turn converts with time into quartz. However, opal-A can pass directly into quartz (Kastner, 1981). According to

Calvert (1974), the transformation of opal-CT into quartz is a solid-solid inversion whose rate depends solely on temperature.

The rate of this chert diagenetic evolution depends on different factors: in situ temperature (Li, 1995; Carson, 1991; Calvert, 1974; Kastner et al., 1977), pore fluid chemistry, pH, time, host rock mineralogy, organic matter abundance, clay minerals, burial depth, host bed porosity and permeability, specific surface area (Williams et al.,

1985; Tribble et al., 1995). Von Rad and Rösch (1974) state, “under normal deep-sea conditions, time appears to be the single most important parameter” (p. 344). The ideal diagenetic sequence derived from the -specific surface area graph of Williams et al. (1985) would be represented by a sedimentary column in which opal-A overlies opal-

CT, that in turn overlies quartz, with transitional zones of mixed phases. Kastner et al.

(1977) demonstrated experimentally the strong influence of the host rock and the composition, along with time and temperature, over the transformation rate of opal-A to opal-CT. Their conclusion was that the transformation rate is greatly enhanced within a carbonate host rock, as compared to clay-rich host sediments, by highly alkaline rich in magnesium. The idea is that magnesium (provided by the seawater) and hydroxide

32

(OH-, provided by dissolution) act as magnesium hydroxide sites of opal-CT

nucleation and subsequent growth of opal-CT lepispheres. The affinity between

carbonate and chertification is explained in terms of alkalinity – by calcite dissolution the

pore water solutions become alkaline favoring the transformation of opal-A into opal-CT at a higher rate. According to the experiments by Kastner et al. (1977), active clay minerals, such as smectites, retard the opal-A to opal-CT transformation, by competing with opal-CT for the available alkalinity in seawater, which results in an enrichment of the clay minerals in Mg.

5.3. Chert deposit types: characteristics and genesis theories

Chert deposits in the geological record are preserved mainly as bedded and nodular cherts. They can form in shallow and deep waters, but many shallow water siliceous deposits do not become cherts because they are masked by a high terrigenous input. Numerous Mesozoic and Cenozoic bedded chert deposits occur in orogenic belts and within ocean floor deposits.

The difference between bedded and nodular chert deposits is regarded as either a morphological difference (Boggs, 1987) or as a genetic difference. Bedded cherts are considered primary accumulations of siliceous tests of planktonic silica-secreting organisms, such as radiolarians for the Paleozoic (radiolarites), and radiolarians and

diatoms (diatomites) for the Mesozoic and Cenozoic (Grunau, 1965; Aubouin, 1965;

Ramsay, 1973; Hein and Parrish, 1987; Maliva and Siever, 1989), whereas nodular cherts

are regarded as diagenetic products resulting from silica remobilization and mineral

33

precipitation within a host rock (Maliva and Siever, 1989; Maliva et al., 1989).

Some authors differentiate between the two types based on the host rock – the chert

deposits occurring in carbonate host rock being considered nodular cherts, and the chert

deposits occurring in siliceous sequences termed bedded cherts (Wise and Weaver,

1974).

5.3.1. Bedded Cherts

General characteristics

Bedded cherts, regarded as primary chert accumulations of either diatom or

radiolarian tests, can form in both deep and shallow water. Presently silica is

concentrated in deep oceans, below the CCD, forming siliceous oozes composed entirely

of biogenic silica of radiolarian tests and diatom . Bedded cherts can form also

beneath shallower waters on the shelf, if surface waters are fertile, there is a paucity of calcareous , and there is a very low terrigenous sediment input. Therefore, the

formation of bedded cherts depends directly on the organic productivity of the planktonic

siliceous organisms (diatoms and radiolarians), that in turn depends on oceanic

circulation and nutrient supply, on the sediment input, the presence of calcareous plankton, and water depth. Aubouin (1965), discussing radiolarites, defined them from two points of view: 1) petrographically and tratonomically. The former form through test accumulation, occurring in all sedimentary environments of an epicontinetal domain, and considered of local significance. The latter are composed of a massive, homogenous sequence of siliceous facies “uniquely characteristic of geosynclinal chains, and

34

particularly of the eugeosynclinal furrows” (p. 115). These can be associated with pelagic

limestones, limestone microbreccias, pelites, , and ophiolites.

Origin and genesis theories

The abiogenic versus biogenic 1origin of chert has long been debated. The main

factors supporting the inorganic origin theory are the association of cherts with volcanic

complexes in many cases, on the presence in certain bedded cherts of some minerals

regarded as volcanic alteration products, such as montmorillonite, palygorskite, sepiolite, and clinoptilolite (Wise and Weaver, 1974), and on the lack of siliceous skeletal remains.

Bedded cherts associated with bentonites and devoid of any siliceous skeletal remains are widely considered to be inorganic in origin.

At the beginning of the twentieth century Davis (1918, in Wise & Weaver, 1974), studying some radiolarian bedded cherts of the Franciscan mélange in California, proposed another theory known as the “gel theory”. This theory explains bedded cherts as being formed by the solidification of layers of silica gel entrapping radiolarian tests as the gel was solidifying. The main mechanism that induces a high accumulation rate of siliceous tests of the planktonic silica-secreting organisms, and ultimately the transformations into bedded cherts, has been debated for centuries: submarine volcanism- induced plankton blooms (the so-called volcanic-sedimentary hypothesis of Aubouin,

1965) based on the spatial association between many bedded cherts and igneous rocks in eugeoclinal settings (Grunau, 1965; Aubouin 1965); abyssal accumulations below the calcite compensation depth, unrelated to igneous activity (Aubouin 1965; Garrison, 1974;

35

Heath, 1974); and upwelling-induced high planktonic productivity (Calvert 1966;

Ramsay, 1973; Hein and Parrish, 1987).

The concept of maturation of siliceous oozes from an initial biogenic silica ooze through porcelanite to chert, was introduced by Bramlette (1946, in Kastner et al., 1977).

In concordance with this theory, the transformation of biogenic opal-A is regarded and interpreted as a process dependent on depth and temperature and ultimately is viewed as a time-dependent process (Wise and Weaver, 1974; von Rad and Rösch, 1974).

5.3.2. Chert nodules

General characteristics

Chert nodules are highly variable in size and shape, ranging from spherical to irregular. They may coalesce to form more continuous layers resembling bedded cherts.

Their internal structure can be laminated or not, and they may contain silicified fossils.

Chert nodules, regarded as diagenetic products, occur predominantly in carbonate rocks such as shelf limestones, pelagic and chalks, in environments ranging from tidal flats to deep ocean basins (Maliva and Siever, 1989). According to Maliva and

Siever (1989; p.425), “there is no intrinsic depositional environmental restriction on nodular chert formation”. Nodular cherts form in hypersaline, marine, and mixed marine- meteoric pore waters. Also, nodular cherts can form in mudrocks, evaporites, burrow fills, but to a much smaller extent, and can nucleate around fossils. In general, chert nodule distribution is a non-random stratigraphic event tending to occur parallel with the bedding planes. The great affinity between carbonate rocks and silica precipitation was in part explained experimentally by Kastner et al. (1977).

36

Origin and genesis theories

Based on mineralogical transformation, Wise and Weaver (1974) discuss 2 basic

theories of nodular chert formation: 1) the maturation theory, and 2) the quartz

precipitation theory. According to the maturation theory proposed by Heath and Moberly

(1971, in Wise and Weander, 1974), chert nodule formation begins with precipitation of disordered cristobalite that in time converts to chalcedonic quartz from the nodule center

toward the periphery.

The quartz precipitation theory was suggested by Lancelot (1973a). To this theory

the driving mechanism of chert nodule formation is primary quartz precipitation, with

disordered cristobalite representing a by-product of the process. The theory is applicable

only for carbonate ooze sequences in which the chert mineralogy is exclusively

represented by quartz. The steps are: precipitation of quartz, growth of the quartz grains

along a quartzification front assuming a high permeability of the sediment, and

precipitation of cristobalite at the periphery of the quartz nodule.

Analyzing the two theories concomitantly with their own analyses of the

Kerguelen chert through scanning electron microscopy, Wise and Weaver (1974)

concluded that chert nodule formation occurs as follows: 1) diffusion of supersaturated

fluid with respect to silica; 2) precipitation of cristobalite lepispheres; 3) development of

a dense nodule nucleus with concomitant dissolution and expulsion of host rock

carbonate; 4) growth of nodule; 5) inversion of the cristobalite to quartz from the center

toward the periphery; 6) pore space filling by fibrous chalcedony and cryptocrystalline

quartz; and 7) growth until the depletion of dissolved silica.

37

Maliva and Siever (1989) describe three earlier models proposed for the origin of

nodular chert, and propose themselves a fourth chert nodule genesis model: 1) the

organic-matter oxidation model, 2) the hydrogen-sulfide oxidation model, 3) the mixing

zone model, and their own model 4) force of crystallization-controlled replacement

model.

The organic-matter oxidation model claims that the decomposition of organic matter may induce early diagenetic nodule formation. When organic matter is oxidized,

the pore water CO2 partial pressure increases driving an increase in the solubility of carbonate minerals. On the other hand the solubility of amorphous silica decreases with decreasing pressure. This model is applicable only for early diagenetic chert nodules, occurring as long as there is organic matter available for oxidation (Siever, 1962).

The hydrogen-sulfide oxidation model was proposed by Clayton (1986, in

Maliva and Siever, 1989) to explain the flint nodule formation in the Upper Cretaceous

Chalk of Western Europe. According to Clayton, the flint nodules formed at the oxic-

anoxic boundaries within the sediment. Below the anoxic-oxic boundaries anaerobic

bacterial sulfate reactions produce hydrogen sulfide (H2S). If the sediment is clay-rich,

the H2S will combine with derived mainly from the clay minerals and will form iron

sulfide. But if the sediment is clay-poor, the excess of H2S will migrate toward the oxic- anoxic interface where it will combine with free molecules of oxygen according to the following reaction:

2- + H2S + 2O2 SO4 + 2H

38

The released hydrogen ions will react energetically with the calcite rock resulting in

intense calcite dissolution. At the site of this boundary the pH registers a significant

decrease allowing the nucleation of opal-CT and formation of flint nodules.

The mixing model was proposed by Knauth (1979) who suggested that many

nodular cherts in limestones have formed in mixed meteoric-marine coastal systems.

According to this model, the mixing zone is characterized by diagenetic solutions that are

undersaturated with respect to calcite and , and supersaturated with respect to

silica. This geochemical environment is one where carbonate can be replaced by silica

under equilibrium conditions. The assumptions of the model are: the meteoric water

passes through a carbonate stack rich in debris of biogenic silica (“proportional to the amount of offshore silica production”, p. 274), a relatively high permeability of this zone, and the system is closed with respect to CO2.

Assessing the three previous models, Maliva and Siever (1989) conclude that

none of them are consistent with their observations obtained by studying six limestone

sequences of different ages (Paleozoic, Mesozioc, and Cenozoic), nor with the different carbonate diagenetic histories. Their model is that calcite dissolution during chertification is restricted to areas of replacement and that the rate of silica precipitation has to be equal to the calcite dissolution rate. None of the previous three models explains these two necessary conditions in the formation of chert nodules and therefore they proposed the

force of crystallization-controlled replacement model. This model is an expression of

the force of crystallization combined with intergranular pressure solution. According to

the model, the undersaturation with respect to calcite occurs only at silica-calcite contacts

39

and therefore calcite dissolution is restricted to sites of silicification. The rate of silica

precipitation is then equal to the rate of calcite dissolution. This model explains the

formation of chert nodules in all environments and suggests that the only necessary condition for chert nodule formation is the presence of pore water supersaturated with respect to opal-CT or quartz. Maliva and Siever (1989) listed three properties of the host sediment as possible controls of chert nodule nucleation and the inter- and intra-bed distribution of chert nodules: 1) organic matter content, 2) porosity and permeability influencing the silica transport and molecular diffusion, and 3) biogenic-opal concentration.

The force of crystallization-controlled replacement model of Maliva and Siever

(1989) explains only how individual crystal grains of quartz or opal-CT can replace

carbonate grains. The formation of a chert nodule consists of silica nucleation and

growth. Nucleation is not enough to generate nodule formation but has to be

accompanied by nodule growth. Any crystal formation includes two steps: 1) the

formation of crystal embryo by nucleation, and 2) the crystal growth. It is known that

opal-CT nucleation during chertification is principally a heterogeneous nucleation on pre-

existing opal-CT crystals (isomineralic heterogenous nucleation; Maliva and Siever,

1989). If a nucleus forms in contact with some surface, commonly another solid, the

nucleation process is called heterogenous nucleation in contrast with the homogenous

nucleation in which nucleus form in the bulk of the solution (Drever, 1997). However,

chert nodule formation is a combined process of heterogenous and homogenous

nucleation of opal-CT lepispheres, followed by crystal growth (Maliva and Siever, 1989).

40

The growth of the silica phase generates a pressure at the silica-carbonate grain contact, driving an increase in the Gibbs free energy and therefore in the solubility of the calcite.

5.4. Time and temperature of chertification and the depth of formation

According to von Rad and Rösch (1974), a period of 70-90 My is necessary for an initial biogenic silica sediment to become a mature quartz chert. Kastner et al. (1977) suggested that siliceous oozes can transform into cherts in 25-50 My. If the in situ temperature is increased by different mechanisms, for example by burial, the period necessary for the completion of the opal-A to quartz transition decreases. For a temperature close to 30ºC, less than 30 My are necessary to form quartz (Heath, 1973, in von Rad and Röch, 1974). However, these estimates are valid only if the silica diagenetic history includes the whole maturation process: opal-A passes into cristobalite, that then passes into quartz. If quartz precipitates directly form solution during early diagenesis as suggested by Lancelot (1973a,b), the age estimates does not apply.

Chert nodules can represent early or late diagenetic events occurring at moderate to deep burial depth (Maliva and Siever, 1989). According to Maliva and Siever, the depth of chert nodule formation can be estimated by the presence of ghosts of intergranular pressure solution contacts and deformed grains. However, they assert that the minimum depth of burial and the time of chertification are uncertain. Studies have showed that opal-CT forms before significant burial at temperatures close to the temperature of the overlying bottom water, while granular microcrystalline quartz forms during deeper burial at higher temperatures (Wise and Weaver, 1974).

41

Data from the Monterey Formation of California suggest an even lower

temperature for chert formation (Matheney and Knauth, 1993). Oxygen isotopic analysis

of opal-CT and quartz suggest that silica diagenesis can occur at temperatures as low as

17º - 21ºC, instead of 35º - 50ºC as previously supposed to be for opal-CT, and 44º –

18 77ºC for quartz, while δ Opore water = 0‰ (Murata et al., 1977; Pisciotto 1981). Based on

the difference in formation temperatures of the opal-CT and quartz chert samples studied

(differences of 40º - 50ºC), Matheney and Knauth (1993) concluded that quartz chert

formed from opal-CT at approximately 900 – 1000 m beneath the opal-CT transition

zone, and that opal-CT formed from biogenic siliceous deposits at approximately 330 m

beneath the sea floor. The assumptions of the model are: a geothermal gradient of

45.2ºC/Km, a sea-floor temperature of 5ºC, and the precipitation temperature of opal-CT

of 20ºC.

42

Chapter 6

METHODOLOGY

6.1. Data acquisition

Compilation of the worldwide Ordovician early diagenetic in situ chert deposits, such as bedded cherts, nodular cherts, and associated silicified oolites yielded a high resolution record at the epoch/series level and, in some cases, at the stage/age level. The compilation is based on several sources: a) online databases - GeoRef, Science Citation

Index Expanded, geology journals, General Science Full Text, and Journal Storage, the scholarly journal archive (JSTOR), b) electronic journal archives such as the American

Association of Petroleum Geologists Bulletin since 1917, and the CD-ROM archive of the Journal of Sedimentary Research between 1931 and 1997 and, c) printed articles. All the electronic sources (GeoRef to a lesser extent) present a predilection toward the North

American literature, therefore a chert database built solely based on them would have presented a strong bias toward the North American continent (Tomescu and Kidder

2002). The bias was highly diminished toward an equitable balance between the North

American continent and the rest of the world, by running Georef literature searches, and especially by analyzing numerous printed reports. The electronic search, using “chert” and “Ordovician” as keywords, yielded around 1900 articles, but only around 400 of those articles were relevant to this study. Approximately 100 articles obtained in addition to those found using electronic search engines produced useful results.

43

6.2. Data management

The chert database was designed, built and managed using FileMaker Pro software. The wide range of information collected is shown in Figure 6.1. The chert occurrences were plotted on Ordovician paleogeographic maps at the epoch level, using the

Paleogeographic Information System/MacTM program, version 5.0, designed by Ross

(1997) based on the data of C.R. Scotese.

6.3. Data structure

Each chert occurrence included in the FileMaker database is characterized by four main types of information: 1) bibliographic references, 2) geographic localization, 3) geologic information and 4) comments (fig. 6.1). The bibliographic references section contains all the useful works analyzed and used in extracting any valuable data pertaining to the chert deposit considered. The geographic localization section includes the name of the area where the outcrop was described (country, region, town) and the geographic coordinates (the latitude and the longitude). The geologic information section comprises information about: a) the type of the chert deposit (nodular versus bedded versus oolitic chert) and its content in siliceous organic remains, if such data was available; b) the name of the rock formation and, if applicable, that of higher rank stratigraphic units; c) the geologic age at the epoch level (Lower/Early, Middle and Upper/Late Ordovician), and even at higher resolution if data was available (e.g., Caradocian, Ashgillian); d) the associated lithologies and the stratigraphic position of the chert unit within the

44

Figure 6.1: FileMaker database entry sample.

45

sedimentary sequence; e) the depositional environment in which the chert deposit was

formed, with the aim to differentiate lagoonal/peritidal from shelf and from slope/basin settings; f) tectonic and paleogeographic settings; and g) diagenetic features of the chert deposit. The workable database built for the present study and that has been used in quantifying the Ordovician chert distribution, includes only those Ordovician chert deposits of well defined age at the epoch level, or as strata bearing siliceous organisms

but not defined as chert deposits.

All these various types of data linked to a chert deposit represent important

information in assessing the determining factors that could lead to the accumulation and preservation of a certain type of chert deposit, in differentiating late from early diagenetic

cherts, in evaluating the origin of a chert deposit, and etc. For example, closed, semi-

open, epeiric, and foreland basins should be regarded differently from the point of view

of chert formation because they are characterized by differences in specific features such

as tectonic activity, volcanism, local sources of dissolved silica, and local oceanic

circulation patterns.

6.4. Data manipulation

The main goals followed in assessing each chert deposit were: a) to establish the

age at the epoch/series level (Lower versus Middle versus Upper Ordovician) and, where

data were available data, at higher resolution, respectively at the stage/age level,

especially for the Upper Ordovician (Caradocian versus Ashgillian); b) to identify the

type of the chert deposit (nodular versus bedded cherts and subsequently silicified

46

oolites); c) to determine if a chert unit contains siliceous organic remains (siliceous spicules and/or radiolarians); d) to recognize the depositional environment in which the chert deposit formed (peritidal/lagoonal versus shelf versus slope/basinal

settings); and e) to infer the genesis of the chert units.

Some of the recorded chert occurrences were obtained from general descriptive

sedimentologic and stratigraphic studies of particular areas and sedimentary sequences of

interest. Many of theses cases bore useful and reliable information pertaining to the cherts

(age, origin, fossil content, and depositional environment). The remaining occurrences

were collected from different geological studies focusing on problems other than the

cherty formations. Most of these merely mentioned the chert occurrences and in these

cases, various interpretations related to the goals of the study were made insofar as the available data allowed.

6.4.1. Interpreting the type of a chert deposit – nodular chert versus bedded

chert

The types of in situ chert deposits considered and quantified herein are: nodular

cherts, bedded cherts, deposits with both nodular and bedded cherts, and “type

undetermined”. A priori, while compiling the data, the difference between bedded and

nodular chert deposits was regarded mainly as a morphological difference, and only

subordinately as a genetic difference. Laminated and stringer chert units have been

considered bedded chert deposits, even though a layer/bed of rock usually refers to a

thickness from a few inches to a few feet. The reason for this decision is that they still

represent a layered sedimentary structure that occurs continuously on a certain surface. In

47

contrast, nodular morphology is a spatially-restricted chert occurence within a host rock.

Many terms are used in the literature to designate a chert deposit. These include: chert,

flint, jasper, ooze, ribbon chert, radiolarian chert, spiculitic chert, chert with radiolarians

and sponges, solid chert, chert in shales, chert lime, chert facies, cherty limestone, cherty dolomite, cherty carbonate, silicified limestone or dolomite, cherty argillite, cherty or chert shale, cherty mudstone, siliceous shales, siliceous bed, siliceous argillites, siliceous

mudstones, and siliceous slate-like shale. The term “siliceous” is ambiguous and

inconsistently used in the literature. For example, theoretically, siliceous shales refer to

shales with very high silica content, as much as 85%, compared to the average shale that

has around 58% silica (Pettijohn, 1975). By extrapolating the definition of the siliceous shales to any siliceous deposit type we should expect to have a sedimentary rock very rich in silica (much above 50%). On the other hand, Shepard (1973) and Duxbury et al.

(2002), in analyzing the deep-sea sediments, define a as a that contains more than 30% biogenous material. Usually, the term “siliceous” is used as a descriptive word for any rock containing noticeable chert components. Hence, in some cases it was hard to discern a compact and well-individualized chert-bearing unit from a siliceous deposit that could bear siliceous particles disseminated within a host rock. Still,

“siliceous” deposits containing siliceous skeletal remains could give valuable information about the silica secreting organisms living in the proximity of that depositional environment. Therefore, if a chert deposit was mentioned as a “siliceous” deposit with no further information regarding any siliceous faunal remains or the chert itself, the deposit was not taken into account as a “true” chert unit, but was still recorded as a putative chert

48

deposit in the expanded chert database. Numerous examples of such deposits were

recorded from China and Kazakhstan.

Many “siliceous” deposits are mentioned in the literature as radiolarian-bearing

strata devoid of further details regarding the type of the chert deposit, depositional

settings, and associated lithologies. Even though they are presumably bedded cherts,

these chert facies were not classified as well-individualized chert deposits such as nodular

and bedded cherts, but as an undetermined type, in order to prevent a supra-

representation of one type of chert deposits over the others.

6.4.2. Interpreting the depositional environment of a chert deposit

Environmental interpretations of the chert occurrences belong to the original

authors, and rely on paleogeography, lithologic associations, and

other paleobathymetric and environmental reconstruction criteria. Based on this

information chert-forming environments were classified into three categories: 1)

lagoonal/peritidal 2) shelf and, 3) slope/basin. Difficulty in determining the environment

of some deposits is an ongoing problem (Murray, 1994), leaving some of the analyzed

cherts in an undecided category (see Appendix).

The general paleogeography of the area in which occurred provided

useful information with respect to the depositional environment of the chert deposits.

Within this context, clastic and carbonate platforms point to shallow water – shelf environments in miogeoclinal settings. Tectonically active settings in which volcanism is associated with clastic sediments and located away from the craton, suggest deeper water settings.

49

Lithologic associations, sedimentary structures and other depositional

environment indicators have represented some of the best tools in evaluating the

depositional setting. A peritidal/lagoonal depositional setting was inferred based on the

presence of paleobathymetric indices represented in many cases by silicified oolites

and/or, to a lesser extent, associations. Cruiziana ichnofacies and hummocky

cross-stratification indicate a shelf depositional setting (Pemberton et al., 1992; Pratt et

al., 1992). Graptolite-rich shales, turbidite facies, flysh sequences, ophiolitic, pillow-

and volcaniclastic chert associations are ascribed to basin/slope environments

(Aubouin, 1965; Grunau, 1965; Murchey et al., 1983). Bentonites, the chemical alteration

product of volcanic ash, might represent significant silica sources due to their high content of clay minerals, covering vast areas. Consequently, some bedded chert deposits have been discovered associated in the geologic record with volcanic ash layers. Owing to the fact that bentonites do not reflect local environmental conditions, bedded cherts associated with them were not interpreted in terms of depositional environment in which they have been found.

6.4.3. Biases in designating a chert deposit type and the corresponding

depositional environment in which it formed

It is tempting to classify any radiolarian-bearing chert as bedded and any -bearing chert as nodular, but it has been demonstrated (Coniglio, 1987, for the

Cow Head Group) that both types of siliceous skeletal remains can occur in both bedded chert and nodular chert deposits. Therefore no such assumptions were made in the absence of other pertinent information.

50

Using the type of the host rock as an indicator for the depositional environment in

which cherts occur, introduces another strong bias. In the case of limestone host rocks,

chert deposits tend to be classified as nodular, shallow water, or shelf deposits. At the other end of the spectrum, if the host rock is shale, especially black shale with graptolitic fauna, the tendency is to designate the chert deposit as bedded, deep water, or basinal

(e.g., the geosynclinal cherts of Pettijohn, 1975). Biogenic components and pelagic clays dominate many deep-sea sediments (Kennett, 1982). However, even though these associations between host rocks and depositional environments are in most of the cases

consistent, some exceptions to the rule are inevitable.

Pettijohn (1975) described bedded cherts, termed cratonic cherts, from shallow

water limestone sequences, which he considered stable shelf associations (p. 401). Also, a

siliceous argillite deposit can accumulate in various settings extending from shallow

water shelves to deep-water environments (Murray 1994; Girty et al., 1996). Murray

(1994) asserts, “general criteria by which depositional environments of cherts can be

determined are not yet available” (p. 213). Coniglio (1987) in describing the chert units

of the Cow Head Group of western Newfoundland, pointed to the presence of chert

nodules in shale sequences. Berry and Wilde (1978) in analyzing lower Paleozoic black

shales concluded that they can form on shelves, slopes and in deep-sea settings. Their

differentiation is possible by integrating paleotectonic and associated lithologic

information. For example, a eugeoclinal tectonic setting correlated with volcanic rocks

indicates deep-sea environments. This tectonic setting could also be a back-arc basin (e.g.

the Sea of Japan) which is of shelf depth. I am not aware of any chert in back-arc basins,

51

perhaps because of siliciclastic dilution. In the absence of other types of evidence, no

depositional environment or chert deposit type was assigned.

6.4.4. The fossil content analysis

No assumptions were made regarding the content of siliceous organic remains.

Therefore, only explicit information from the analyzed works has been taken into

consideration in assessing the siliceous fossil content of the rocks.

6.4.5. The geologic age adjustment

A crucial requirement for compiling worldwide chert occurrences with epoch-

level resolution is the integration of different Ordovician stratigraphic scales – British,

Scandinavian, Russian, North American, Chinese, Australian (fig. 6.2.). In order to compare and interpret worldwide chert deposits within a consistent temporal framework at the epoch/series level, and even at a higher temporal resolution, a unifying correlative

Ordovician stratigraphic scale was built and used (based on Nikitin et al., 1986; Harland

et al., 1990; Webby, 1995; Harris et al., 1995; Palmer and Geissman, 1999; Veevers,

2000; Ferguson and Fanning, 2002). The Ordovician time scale used herein as a template

is the “1999 ” of the Geological Society of America (Palmer and

Geissman, 1999) that in turn relies on the British stratigraphic nomenclature.

Potential for confusion arises at the stage level in the Upper Ordovician. This epoch includes the Caradocian and Ashgillian stages. Unfortunately, the epoch placement of the Caradocian stage has not been consistent through time. Presently it is widely

52

Figure 6.2: Ordovician correlation chart used in this study (compiled from Nikitin et al., 1986, 1991; Harland et al., 1990; Harris et al., 1995; Webby, 1995; Palmer and Geissman, 1999; Veevers, 2000; Ferguson and Fanning, 2002).

53

accepted that the Caradocian represents the basal stage of the Upper Ordovician epoch.

However, until less than one decade ago, the American geological literature considered

the Caradocian stage as the uppermost stage of the Middle Ordovician. Even some of the

most recent works that deal with the Ordovician use this stratigraphic system. One of the

main problems encountered during this study in terms of age, was facing published

Ordovician reports that did not define the stratigraphic system used. Hence, a late Middle

Ordovician chert deposit in the old stratigraphic system is an early Upper Ordovician deposit according to the new stratigraphic system. In concordance with the newest geochronological scale for the Ordovician period (Palmer and Geissman, 1999), the duration of the Caradocian stage was 9 My (458 Ma – 449 Ma). Including these 9 million years into the Middle or the Upper Ordovician changes significantly the duration of the two epochs. Furthermore, analysis and interpretation of the abundance of the same rock type at a worldwide scale for periods of time with coincident names but different age ranges can yield considerable errors.

6.4.6. The chert occurrence quantification guidelines – over-representation

versus under-representation of the Ordovician chert abundance at the epoch

level

The basic geological stratigraphic unit, that is, the formation, represents the quantifying unit used in this study: one chert-bearing formation is counted as one chert occurrence. At the epoch level, the same one chert-bearing formation spanning over two

Ordovician epochs or even the whole Ordovician, can result theoretically in two and respectively three chert occurrences. Considering any deposit of suspicious age as a

54 deposit that spans multiple epochs, and counting it as such, could eliminate the errors, but at the same time it would generate a over-representation of chert. The risk of such a chert over-representation is even greater if we consider that one of the most problematic

Ordovician age limits at epoch level is that between the Middle and the Upper

Ordovician, a time demonstrated herein as a transition from a period poor in chert formation to a very rich chert deposition period. In view of these facts, the chert deposits of uncertain or widely spread age were eliminated, and only those with well defined epoch or stage ages were considered in the quantitative results, even though this decision results in an under-estimation of the worldwide Ordovician chert abundance.

Each formation was considered as only one chert occurrence, even for formations with more than one chert members. If a group had more than one chert formation, as many chert occurrences as defined formations were taken into account. Additionally, the same formation can correspond to two different adjacent depositional environments

(lagoonal/peritidal and shelf or shelf and slope/basinal settings), in which case it resulted in two chert occurrences, each corresponding to one of the two depositional settings. The chert deposits associated with bentonites were not counted in the quantification phase of the depositional environment analysis, and are not included in chert abundance plots.

They were nevertheless recorded as bedded chert deposits.

6.5. Pitfalls/Problems of a study compiling data from published reports:

In any compilation of data from multiple sources, inconsistencies arise. Some of the problems encountered in this study are listed below.

55

1) Older rocks are less represented in the rock record and therefore an

extrapolation of the Ordovician chert deposits with the aim to compare their abundance

with those of other geological periods could generate errors. Normalization to outcrop area can lessen the impact of this problem at the system level, but to date outcrop area has

not been calculated at the epoch and stage level for the Ordovician.

2) Some chert deposits are less well documented than others, resulting in sparse

information regarding the age, , sedimentology, faunal content, depositional environments, and etc.

3) The lack of information of the published data relative to the definition of

certain geological notions whose connotations may largely vary, with strong implications

on interpretation (e.g., the Ordovician time scale used considering that it has been

changed through time, the common shallow- vs. deep-water setting ambiguity with

respect to the basin morphology and implicitly the basin type).

4) Some stratigraphic intervals are easier to date than others depending mainly on

the type of faunal content.

5) Some paleoenvironmental settings are characterized by particular factors (e.g.,

starved, closed basins) and therefore certain inter-basinal correlations are pointless.

Moreover, the formation of chert deposits within such localized-conditions may not

follow a general chert genesis mechanism, but they are useful because they evidence

endemic conditions and the amplitude (at the basin and even global level) of a peculiar

mechanism responsible for their formation.

56

6) Stratigraphic inconsistencies regarding definitions of formations and groups

lead to some artificial variations in chert abundance. For example, an over-estimation of

the chert occurrences is possible when a sedimentary unit with multiple chert horizons

was considered a group and not a formation (e.g., the Plattin Group with three chert

formations resulting in three chert occurrences; before it was considered a formation and

not a group, and would have been recorded as only one chert occurrence).

7) The paleogeographic position of some micro-continents is still debated and thus the data interpretation with respect to the paleolatitudinal distribution of the occurrences encounters errors. An eloquent example is that represented by the Argentine

Precordillera which is considered either as part a of Gondwana that collided with

Laurentia during the Ordovician (Dalla Salda et al., 1992), or as a microcontinent that rifted from the Ouachita embayment of North American Laurentia during the Cambrian and collided with Gondwana during the Ordovician (Thomas and Astini, 1996).

8) The worldwide rock sequences are unequally studied due to objective impediments (relief, scientific potential, political regimes, and etc.), and thus the absence of a geological character of certain geographical areas might reflect a lack of knowledge and not a physical absence. A prime and very strong bias of a worldwide compilation study is represented by the geographic component.

9) The published information easily accessible is strongly biased toward the North

American continent, and thus geological studies achieved in regions with a weak

international communication system are highly underrepresented in databases (e.g., South

America, Western Europe, Asia, Eastern Europe, and etc.). This bias is even larger if we

57

consider that these modern regions were representing different paleocontinents,

paleolatitudes, and very likely different physical and biologic ecosystems.

6.6. The usefulness of a study compiling data from published reports:

1) This study integrates the biotic component with the abiotic physico-chemical

factors as interconnected parts of the same Earth system.

2) Lithologic compilations offer a general image of the worldwide Ordovician early diagenetic chert occurrences such as bedded, nodular cherts, and silicified oolites.

3) Paleogeographic chert distribution can be used to infer paleo-upwelling areas and/or high endogenous and exogenous dissolved silica and other nutrient inputs that could lead to the organic, inorganic or organic but induced inorganically chert formation.

The study can also offer a good image of the Ordovician shelf areas suggesting the spatial position of the paleo-shorelines.

4) Based on depositional settings, siliceous organism migration patterns, as well as potential factors responsible for the formation, accumulation, and preservation of chert units can be inferred.

5) Local and general oceanic circulation patterns can be distinguished pointing toward local and global events.

6) Possible source rocks and hydrocarbon reservoirs could be identified.

7) The study offers the possibility of not only intra-basinal correlations, but also of inter-basinal stratigraphic correlations.

58

Chapter 7

RESULTS

The comprehensive compilation of Ordovician cherts yielded 191 deposits covering 5 continents: , South America, Europe, Asia, and . The appendix includes all chert formations that have been found, regardless of the degree of certainty of their age. However, the analysis of chert abundance includes only the 161 chert deposits with ages well defined at the epoch level or higher resolution. The descriptive information included in the appendix represents a concise statement of the information in the FileMaker database that is available upon request.

7.1. Epoch - level temporal trend in the chert abundance

Figure 7.1A illustrates the epoch-level chert distribution expressed as both raw data (counts) and time-normalized data (Table 7.1). Both representations show the same pattern throughout the three Ordovician epochs: abundance peaks for the Lower and

Upper Ordovician, contrasting with a sharp decline during the Middle Ordovician. More than half of the Upper Ordovician chert deposits (approximately 51%) are Caradocian, 14 deposits are attested as Ashgillian (19%), and the rest of 22 deposits (30%), have undifferentiated Upper Ordovician age (fig. 7.1B). Time normalized data (Table 7.1) shows that the Caradocian chert peak stands as the highest Ordovician chert abundance

59

Table 7.1: Temporal distribution of the chert deposits at epoch and age/stage level. Raw and time-normalized data.

Normalized data to Duration Epoch No. No. time N/Ma Ma

Ashgillian 14 2.3 6 Upper 74* 4.9 Caradocian 38 4.2 9

Middle 24 2 12

Lower 63 3.2 20

Total = 161

* the number represents 14 (Ashillian) + 38 (Caradocian) + 22 (undifferentiated Upper Ordovician)

Figure 7.1: Ordovician distribution of chert deposits. Numbers next to dots represent number of cherts per million years. (A) Epoch level distribution: raw (N) and time- normalized data (N/Ma). (B) Stage level distribution expressed as percentages from a total of 74 chert deposits.

60

interval, with 4.2 chert deposits per 1 million years. If all of the undifferentiated Upper

Ordovician cherts would prove to be Ashgillian, then the maximum chert abundance

would shift toward the Ashgillian. However, this is not likely, because stratigraphic

information on these cherts suggests that some of them may have extended over both the

Caradocian and Ashgillian.

In concordance with the Late Ordovician short duration glaciation theory

(Brenchley et al., 1994), the two stages of the Upper Ordovician represent geological

periods of contrasting climates: the Caradocian corresponds to an ice free world whereas

the Ashgillian was an icehouse world. Hence, the differentiation of the Upper Ordovician cherts at the stage level proves is significant.

7.2. Epoch - level temporal trend in chert types

The chert-bearing formations were differentiated into: i) nodular cherts, ii) bedded

cherts, and iii) formations with both nodular and bedded cherts (Table 7.2). Figure 7.2A

illustrates the epoch-level distribution in both raw and time-normalized absolute counts

for the three types of chert deposits, and shows a trend similar to that of the chert

abundance at the epoch-level: Lower and Upper Ordovician peaks – the latter steeper

than the former – separated by a Middle Ordovician drop. Nodular cherts decline in

relative abundance from the Lower to the Upper Ordovician while the number of bedded

cherts increases (fig. 7.2B), revealing a significant facies change in chert type through the

Ordovician.

61

Table 7.2: Temporal distribution of the chert deposit types at epoch (A) and stage (B) level. Raw, time-normalized, and relative abundance data.. 17* = 3 + 10 + 4 (undifferentiated Upper Ordovician); 28* = 2 + 5 + 11 (undifferentiated Upper Ordovician); 16* = 6 + 7 + 3 (undifferentiated Upper Ordovician).

A Nodular cherts Bedded cherts Nodular & bedded cherts Undetermined Total N Epoch N % N/Ma N % N/Ma N % N/Ma N % N/Ma

Upper 17 23 1.1 28 37.8 1.9 16 21.6 1.1 13 17.6 0.9 74

Middle 4 16.7 0.3 9 37.5 0.8 6 25 0.5 5 20.8 0.4 24

Lower 21 33.3 1.1 20 31.7 1 12 19 0.6 10 15.9 0.5 63

161

B Nodular cherts Bedded cherts Nodular & bedded

Epoch Stage N N/Ma N N/Ma N N/Ma

Ashgillian 3 0.5 2 0.3 6 1 Upper 17* 1.1 28* 1.9 16* 1 Caradocian 10 1.1 15 1.6 7 0.8

Middle 2 0.2 5 0.4 12 1

Lower 21 1.1 21 1.1 17 0.9

62

Figure 7.2A: Temporal distribution of chert deposit types at the epoch level. Raw (N) and time-normalized (N/Ma) data. Numbers next to dots represent number of cherts per million years.

Figure 7.2B: Temporal variation of the relative abundance of chert deposit types at the epoch level (light gray - nodular cherts; medium gray - bedded cherts; dark gray - deposits with nodular and bedded cherts; white - undefined chert deposits).

63

7.3. The spatio-temporal distribution of chert occurrences

Table 7.3 and Figure 7.3 display the variation in time, space and depositional

environment of both raw and normalized count data through the Ordovician. The

variation of the normalized data parallels that of the raw data. The main Ordovician chert

deposition sites are plotted on paleogeographic maps (fig. 7.4, 7.5, and 7.6). The epoch-

level distribution of these sites along with their depositional environments reveal

temporal, spatial, and evolutionary patterns in silica accumulation throughout the

Ordovician.

7.3.1. Lower Ordovician

Chert depositional environments during the Lower Ordovician are dominantly

shallow water carbonate settings in which most silica burial occurred as peritidal

silicified oolites, and as nodular and bedded cherts that are evenly dispersed between

peritidal/lagoonal and shelf environments. All peritidal/lagoonal chert occurrences are

silicified oolite-bearing formations. On the other hand, the slope/basin chert-bearing

sequences (a quarter of the Lower Ordovician total) indicate that deep-water silica

deposition represented a quite significant component of silica burial during this time as

well. The deep water chert facies are dominated by bedded cherts supplemented by nodular cherts (Table 7.4, fig. 7.7A).

The paleolatitudinal range of the Lower Ordovician cherts (fig. 7.4) extends from the equator up to 45 – 50˚ north and south. The deep-water cherts, confined principally to lower paleolatitudes (0 - 30˚), cluster on the southeastern margin of Laurentia (in modern

Canada and Great Britain), plus a few on the Kazakhstan block and Gondwana

64

Table 7.3: Temporal (A epoch level; B stage level) and spatial (Onshore – Offshore) distribution of the chert deposits. Raw and time-normalized data. 33* = 17 + 9 + 7 (undifferentiated Upper Ordovician); 38* = 19 + 4 + 15 (undifferentiated Upper Ordovician); * the total 171 is different from the total 161 (Tab. 7.1) because a couple of chert deposits have been assigned to multiple depositional environments. Raw, relative abundance, and time-normalized data.

A Peritidal/Lagoonal Shelf Slope/Basin Undecided Total N*

Epoch N % N/Ma N % N/Ma N % N/Ma N % N/Ma N %

Upper 3 3.8 0.2 33 42.3 2.2 38 48.7 2.5 4 5.1 0.3 78* 100

Middle 2 7.7 0.2 5 19.2 0.4 12 46.2 1.0 7 26.9 0.6 26* 100

Lower 21 31.3 1.1 21 31.3 1.1 17 25.4 0.9 8 11.9 0.4 67* 100

171*

B Peritidal/Lagoonal Shelf Slope/Basinal

Epoch N N/Ma N N/Ma N N/Ma

Ashgillian 0 0 9 1.5 4 0.7 Upper 3 0.2 33* 2.2 38* 2.5 Caradocian 3 0.2 17 1.9 19 2.1

Middle 2 0.2 5 0.4 12 1

Lower 21 1.1 21 1.1 17 0.9

65

Figure 7.3: Ordovician spatio-temporal distribution of chert occurrences (percentages and time-normalized data). The dots represent number of cherts per million years.

66

Figure 7.4: Paleogeographic distribution of Lower Ordovician cherts

67

Figure 7.4.: Paleogeographic distribution of Lower Ordovician cherts. Red bullets

represent peritidal/lagoonal cherts; black bullets represent shelf cherts; stars represent

slope/basinal cherts; dotted bullets are cherts bearing siliceous sponge spicules from

Argentina, not included in the Appendix; diamonds represent cherts whose depositional

environment is unknown. Solid arrows represent cold currents and dashed arrows

represent warm currents (after Berry et al., 1995). The paleoegeography was obtained using the PGIS/MacTM software, version 5.0 (Ross, 1997)

68

Figure7.5: Paleogeographic distribution of Middle Ordovician cherts

69

Figure 7.5.: Paleogeographic distribution of Middle Ordovician cherts. Red bullets represent peritidal/lagoonal cherts; black bullets represent shelf cherts; stars represent slope/basinal cherts; diamonds represent cherts whose depositional environment is unknown. Solid arrows represent cold currents and dashed arrows represent warm currents (after Berry et al., 1995). The paleoegeography was obtained using the

PGIS/MacTM software, version 5.0 (Ross, 1997).

70

Figure 7.6: Paleogeographic distribution of Upper Ordovician cherts

71

Figure 7.6.: Paleogeographic distribution of Upper Ordovician cherts. Red bullets represent peritidal/lagoonal cherts; black bullets represent shelf cherts; stars represent slope/basinal cherts; diamonds represent cherts whose depositional environment is unknown. Solid arrows represent cold currents and dashed arrows represent warm currents (after Berry et al., 1995). The paleoegeography was obtained using the

PGIS/MacTM software, version 5.0 (Ross, 1997).

72

Table 7.4: Distribution of the chert deposit types by epoch and depositional environment (TL – TM – TU = total Lower – total Middle – total Upper; e.g. 21 - 2 - 3). Raw data.

Peritidal / Lagoonal Shelf Slope/Basin Undefined Epoch Total N B N&B U N B N&B U N B N&B U N B N&BU

Upper 0 0 2 1 16 1 14 2 1 24 4 9 0 2 0 2 78

Middle 1 0 1 0 2 0 2 1 0 7 3 2 2 2 1 2 26

Lower 13 0 7 1 6 5 4 6 0 12 3 2 3 3 1 1 67 21-2-3 21-5-33 17-12-38 8-7-4 171

73

Figure 7.7: Relationships between chert types and depositional environments over time. Relative abundance data. (A) Lower; (B) Middle; (C) Upper Ordovician.

74

supercontinent (in Australia and New Zealand). Only one slope/basin chert deposit occurs beyond 30˚ south on Gondwana (Bolivia). The mid-latitudinal cherts (30˚ - 50˚), which are represented mainly by shelf deposits, are located north of the equator on Gondwana

(China and Korea). All the peritidal/lagoonal cherts are restricted to Laurentia (the south- southwestern margin) at low paleolatitudes, all occur in the USA, except for one in Great

Britain.

Overall, the trend of the silica burial during the Lower Ordovician is dominated by low paleolatitudinal cherts (90.5%), followed by mid-latitudinal cherts (9.5%;Table

7.5). Geographically, approximately 75% of the total Lower Ordovician cherts occur on

Laurentia, averaging 2.4 deposits per My (Table 7.6).

7.3.2. Middle Ordovician

During the Middle Ordovician, a pronounced decline in silica burial occurred, coeval with a significant migration of chert deposits from shallow settings, peritidal/lagoonal and respectively shelf, toward slope/basin sites (fig. 7.3). Middle

Ordovician deep-water environments represent the dominant locus in silica accumulation, mainly as bedded cherts (fig. 7.7B, Table 7.4). The formation of silicified oolites almost ceases (Table. 7.7), and that of the shelf cherts strongly diminishes. A slight decrease is recorded in the abundance of mid-latitudinal cherts concomitant with an increase of the low-latitudinal ones (Table 7.5). Paleogeographically, the Laurentia data reveal a decrease in the chert abundance by approximately 20%, or by more than a half considering the time-normalized data (Table 7.6), most of the occurrences again on the southern margin (fig. 7.5). Only the Kazakhstan block registers a slight increase in the

75

Table 7.5: Temporal trends in the paleolatitudinal distribution of Ordovician silica burial.. Raw and relative abundance data.

0 - 300 30 - 600 > 600 Total N Epoch N % N % N % N %

Upper 71 96 1 1.3 2 2.7 74 100

Middle 22 91.7 2 8.3 0 0 24 100

Lower 57 90.5 6 9.5 0 0 63 100

161

chert abundance. The geographic distribution of deep-water cherts follows the same patterns as that characteristic for the Lower Ordovician ones, that is, the southeastern margin of Laurentia, the Kazakhstan microcontinent, and Australia.

7.3.3. Upper Ordovician

The Upper Ordovician silica burial as recorded by numbers of chert deposits surpasses by far the two older Ordovician epochs (fig. 7.3). Both shelf and slope/basinal deposits record a significant increase compared to both previous epochs, Lower and

Middle Ordovician. A balance between shelf and deep-water cherts, that was tipped heavily toward deep-water cherts in the mid-Ordovician, is restored. The few peritidal/lagoonal chert facies hosting silicified oolites all occur in the Caradocian (Table

7.6). The first high latitude and north margin Laurentian cherts appeared in this time

76

Table 7.6: The temporal and geographical distribution of the chert deposits. Raw, relative abundance, and time-normalized data.

Gondwana & close Laurentia Kazakhstan block Others blocks Epoch Total Duration Ma N N/Ma % N N/Ma % N N/Ma % N N/Ma %

9 1.5 64.3 3 0.2 21.4 2 0.1 14.3 * * * 14 Ashillian: 6 Upper* 28 3.1 73.7 9 1 23.7 1 0.1 2.6 * * * 38 Caradocian: 9

Middle 13 1.1 54.2 5 0.4 20.8 4 0.3 16.7 2 0.2 8.3 24 12

Lower 47 2.4 74.6 13 0.7 20.6 3 0.2 4.8 * * * 63 20

* The Upper Ordovician cherts do not include the deposits whose age is either undifferentiated Upper Ordovician (12 deposits) or span the whole Upper Ordovician (10 deposits). The numbers were obtained from Appendix.

Table 7.7: Temporal distribution of Ordovician silicified oolites. Raw and time-normalized data.

Epoch Stage N N/Ma Ma

Ashgillian 0 0 6 Upper Caradocian 3 0.3 9

Middle 2 0.16 12

Lower 12 0.6 20

77

interval (fig. 7.6; Table 7.5). The low latitude deposits continued to dominate the system.

Deep-basin deposits center around Laurentia and numerous Gondwanan deep chert facies

are identified in Australia. Most of the shelf deposits are limited to the southern and

western margins of Laurasia. The examination of the spatial distribution of the cherts by

continents shows that Caradocian cherts increased by almost a factor of three in

Laurentia, and doubled in Gondwana and the proximal microcontinents (Table 7.6).

7.4. Chert type - depositional environment relationship over time

Figure 7.7 illustrates the relationship between the chert deposit types and the depositional environment in which they formed (Table 7.4). Nodular cherts dominate the peritidal/lagoonal and shelf settings, in contrast to the bedded cherts that dominate deep- water habitats. This relationship between chert type and depositional environment persists throughout the Ordovician: shelf settings are characterized by evenly dispersed nodular and bedded cherts within a carbonate environment, whereas deep-water settings are clearly dominated by bedded cherts within a siliciclastic environment. Nodular cherts can nonetheless occur in slope/basinal settings, and bedded cherts in shelf and even shallower environments. By the Middle Ordovician the silicified oolites almost disappear from the chert system.

78

7.5. The temporal and spatial distribution of the Ordovician silica secreting

organisms present in chert deposits

More than a third of the chert occurrences (37.2%) contain siliceous skeletal remains (Table 7.8). The majority host radiolarians, others siliceous sponge fragments, and a few contain both types (fig. 7.8).

A look at the depositional environments of cherts bearing siliceous organisms reveals that during the Early Ordovician both radiolarians and siliceous sponges inhabited deep as well as shallow water niches (Table 7.9Ab). During the Lower Ordovician most of the radiolarian remains are hosted by slope/basin bedded cherts. The siliceous sponge remains have been found only in nodular and bedded chert type. Most sponge-bearing cherts are found in shallow-water sedimentary sequences (peritidal/lagoonal and shelf), and only two, or possibly three, slope/basin cherts associated with ophiolitic successions preserve siliceous sponges (e.g., the Burubaital Formation in Kazakhstan, the lower part of the Cow Head Group of Newfoundland, and possibly the Durness Limestone of

Scotland). Overall, siliceous sponges are associated with carbonate rocks, and radiolarians with clastic sediments such as shales, graptolitic black shales, and ophiolitic complexes.

During the Middle Ordovician, more than half of the chert occurrences contain siliceous skeletal remains (Table 7.8), the majority represented by radiolarians.

Radiolarian-bearing formations are exclusively bedded chert from deep-water sequences

(Table 7.9 Ba and Bb). The few siliceous sponge-bearing formations are both shallow and deep water cherts of bedded and nodular types.

79

Table 7.8: Epoch level distribution of chert deposits with siliceous skeletal remains. *SSR = siliceous skeletal remains. Raw and relative abundance data.

Radiolarians Sponges Radiolarians & Sponges Total fm. with Epoch % Total SSR* N % % N % N Upper 19 25.7 5 6.8 4 5.4 28 37.8 74

Middle 10 41.7 2 8.3 2 8.3 14 58.3 24

Lower 8 12.7 6 9.5 4 6.3 18 28.6 63

Total 37 23 13 8 10 6.2 60 37.2 161

Figure 7.8: Epoch-level variation of the relative abundance of chert formations that bear siliceous skeletal remains (deposits with radiolarians, deposits with siliceous sponges, and deposits with both radiolarians and siliceous sponges).

80

Table 7.9: Relationships between chert deposit types and types of siliceous skeletal remains for the Lower (Aa), Middle (Ba), and Upper Ordovician (Ca). Relationships between depositional environments and types of siliceous skeletal remain for the Lower (Ab), Middle (Bb), and Upper Ordovician (Cb). Absolute counts.

Lower Ordovician Aa N B N&B U Total

Radiolarians 6 2 8 Sponges 2 2 2 6 Both 1 2 1 4

Total Ab 18 P/L Shelf S/B U Total

Radiolarians 2 7 9 Sponges 3 3 6 Both 1 2 3

Middle Ordovician Ba N B N&B U Total

Radiolarians 6 1 3 10 Sponges 2 2 Both 1 1 2

Total Bb 14 P/L Shelf S/B U Total

Radiolarians 8 2 10 Sponges 1/2 1 1/2 2 Both 2 2

81

Table 7.9: continued.

Upper Ordovician Ca N B N&B U Total

Radiolarians 12 7 19 Sponges 3 2 5 Both 2 2 4

Total Cb 28 P/L Shelf S/B U Total

Radiolarians 16 3 19 Sponges 5 5 Both 1 3 4

Siliceous skeletal remains (SSR), mostly radiolarians, occur in 28 Upper

Ordovician chert facies. Here too, radiolarians characterize deep-water sedimentary

successions, whereas siliceous sponges are found in nodular and bedded cherts in either

shelf or slope/basin settings.

The Ordovician formations with siliceous skeletal remains reveals a clear affinity between radiolarians and deep-water bedded cherts, in contrast to the siliceous sponge remains that can occur in any type of chert deposit (bedded or nodular) and depositional

environment (deep or shallow water) (Table 7.9). From the Lower to the Upper

Ordovician, formations bearing siliceous organisms change from mainly

82

peritidal/lagoonal settings to shelf/basin settings, so that by the Upper Ordovician there is

no evidence of peritidal/lagoonal chert formations bearing siliceous sponge skeletal

remains. Radiolarian-bearing formations also exhibit trend from shelf to slope/basin

settings by the Middle Ordovician, so that in the Middle and Upper Ordovician radiolarians are almost exclusively preserved in deep-water sedimentary sequences.

Most nodular cherts are hosted by carbonate rocks such as limestones and dolomites, and most bedded cherts by shales. This finding supports the view according to which some authors make a difference between the two types based solely on the host

rock – the chert deposits occurring in carbonate host rocks being considered nodular

cherts, and the chert deposits occurring in shale sequences being considered bedded

cherts (Wise and Weaver, 1974).

7.6. Testing the significance of associations between the type of siliceous

skeletal remains, chert deposit styles, and depositional environments

The chi-square test is a statistic used for testing the significance of certain

distributions (the distribution between the columns and the rows of a table) or in other

words for comparing observed data with data expected to be obtained according to a

specific hypothesis. The null hypothesis (Ho) states that there is no significant difference

between the expected and observed data. If any expected value of any category is less

than 5, a correction has to be applied. In this respect, Yates’ correction represents an

arbitrary adjustment to the chi-square test when applied to tables with one or more cells

83

with frequencies less than 5 (e.g., the expected values of the R-P/L and SS-P cells in fig.

7.10). However it can be applied only to 2x2 tables.

The Chi-square test was applied to the following distributions: 1) chert deposit

type (nodular vs. bedded cherts) and silica secreting organisms (radiolarians vs. siliceous

sponges) (fig. 7.9); 2) the silica secreting organisms (radiolarians vs. siliceous sponges)

and the depositional environment (shallow represented by peritidal/lagoonal together

with shelf vs. slope/basinal) (fig. 7.11); and 3) the chert deposit type (nodular vs. bedded cherts) and the depositional environment (peritidal/lagoonal vs. shelf vs. slope/basinal)

(fig. 7.10). All of the distributions proved to characterize statistically significant associations between the tested pairs of variables.

7.7. Evaluation of the efficiency of the method

The global compilation of Ordovician chert occurrences undertaken in this study has several characteristics that single it out among other such attempts, both in its approach, and the results generated. The approach which targeted only one geological period, the Ordovician, allowed extensive data gathering and detailed processing of these data with respect to multiple aspects: the chert deposit type, depositional environment, fossil content with respect to silica secreting organisms, and age. The results thus obtain offer a solid empirical base for tackling several questions related to chert formation and preservation, and allow outlining and discussion of patterns in the spatial and temporal biogenic chert distribution.

84

Figure 7.9: Chi square test of the relationship between the type of siliceous organisms (R = radiolarians; SS = siliceous sponges) and the chert deposit style (B = bedded cherts; N = nodular cherts).

R SS Total

B 34 15 49 N 6 16 22 40 31 71

R - B R - N SS - B SS - N

Upper 16 2 6 7 Middle 9 2 4 3 Lower 9 2 5 6 34 6 15 16

O E O - E (O - E)2 (O - E)2/E

R-B 34 27.61 6.39 40.89 1.48 R-N 6 12.39 -6.39 40.89 3.30 SS-B 15 21.39 -6.39 40.89 1.91 SS-N 16 9.61 6.39 40.89 4.26 10.95

χ calc = 10.94; χ 0.05,1 = 3.84

H0: there is no relationship between the chert deposit type and the organisms H1: there is a relationship between the chert deposit type and the organisms calculated χ > theoretic χ → the null hypothesis is rejected

85

Figure 7.10: Chi square test of the relationship between the siliceous organisms (R = radiolarians; SS = siliceous sponges) and the depositional environment (P/L = peritidal/lagoonal; Sh = shelf; S/B = slope/basinal).

I. The depositional environment is divided in three different depth-related settings

R SS Total

P/L 1 5 6 Sh 3 11 14 S/B 38 8 46 42 24 66

R-P/L R-Sh R-S/B SS-P/L SS-Sh SS-S/B

Upper 0 1 19 0 6 3 Middle 0 0 10 1 1 3 Lower 1 2 9 4 4 2 1 3 38 5 11 8

O E O-E (O-E)2 (O-E)2/E

R-P/L 1 3.82 -2.82 7.94 2.08 R-Sh 3 8.91 -5.91 34.92 3.92 R-S/B 38 29.27 8.73 76.17 2.60 SS-P 5 2.18 2.82 7.94 3.64 SS-Sh 11 5.09 5.91 34.92 6.86 SS-S/B 8 16.73 -8.73 76.17 4.55 23.65

Because three cells have values equal or less than five, and because the Yates’s correction can be applied to only 2x2 table, the significance has been tested between the silica secreting organisms and shallow waters represented by peritidal/lagoonal together with shelf environments vs. slope/basinal settings.

86

Figure 7.10: continued.

II. The depositional environment is divided in two different depth-related settings: shallow (peritidal/lagoonal and shelf) vs. deep (slope/basinal)

R SS Total

P/L & Sh 4 16 20 S/B 38 8 46 42 24 66

O E (O-E)-0.5 (O-E-0.5)2 (O-E-0.5)2/E

P/L&Sh-R 4 12.73 -9.23 85.14 6.69 S/B-R 38 29.27 8.23 67.69 2.31 P/L&Sh-SS 16 7.27 8.23 67.69 9.31 S/B-SS 8 16.73 -9.23 85.14 5.09 23.40

χ calc = 23.40

χ 0.05,1 = 3.84

H0: there is no relationship between the depositional environment type and the siliceous organisms H1: there is a relationship between the chert deposit type and the siliceous organisms calculated χ > theoretic χ → the null hypothesis is rejected

87

Figure 7.11: Chi square test of the relationship between the depositional environment (shallow waters represented by P/L = peritidal/lagoonal together with Sh = shelf vs. deep waters represented by S/B = slope/basinal) and the chert deposit style (B = bedded; N = nodular).

N B Total

P/L&Sh 68 36 104 S/B 11 53 64 79 89 168

O E O - E (O - E)2 (O - E)2/E

P/L&Sh - N 68 48.90 19.10 364.63 7.46 S/B - N 11 30.10 -19.10 364.63 12.12 P/L&Sh - B 36 55.10 -19.10 364.63 6.62 S/B - B 53 33.90 19.10 364.63 10.75 36.94

χ calc = 36.94

χ 0.05,1 = 3.84

H0: there is no relationship between the depositional environment and the type of the chert deposit H1: there is a relationship between the depositional environment and the type of the chert deposit calculated χ > theoretical χ → the null hypothesis is rejected

88

This global compilation resulted in the extensive database of 191 worldwide

Ordovician chert occurrences. Another important contribution of the study resides in the

fact that, in contrast to previous surveys that have focused only on bedded cherts or

silicified oolites, the present study compiles chert occurrences including three types of

chert deposits - bedded cherts, nodular cherts, and silicified oolites. This differentiation provides a strong quantitative foundation for reliable inferences that can led to much deeper, broader, and more detailed insights into various aspects of the silica cycle.

The epoch-level age resolution of this study is the highest achieved for an analysis of chert occurrences. A total of 161 are dated at the epoch level or even higher resolution, at the stage level. Work at this resolution demonstrates that perturbations in the silica cycle exist, they can be perceived at a level of detail higher than the system level, and they can be interpreted in terms of physico-chemical oscillations in the Earth system.

The temporal resolution of the study highlighted problems with stratigraphic nomenclature in previous surveys. Resolution of these problems and compilation of data for the Caradocian and Ashgillian separately was particularly important as these two

Upper Ordovician stages were very different climatically. Their separate treatment reveals important changes in chert abundance between the two stages, changes that would have been assigned to the Middle – Upper Ordovician transition otherwise, and particularly linked to the Late Ordovician glaciation. These observations allows detailed insights into the climate fluctuations during the Upper Ordovician, fluctuations that a gross epoch-level treatment would have overlooked because of lack of resolution.

89

The structure of the dataset provided a comprehensive, and at the same time

detailed, framework for systematic surveys undertaken within this compilation on the

depositional environment of formation corresponding to each chert occurrence

(peritidal/lagoonal vs. shelf vs. slope/basin), and on the type of siliceous skeletal remains

(radiolarians vs. siliceous sponges) preserved in each chert deposit. The volume and

resolution of data thus developed allow quantitative analyses that demonstrat for the first

time, with statistically significant results, relationships between chert deposit types,

depositional environments, and silica secreting organisms. These tests support the organic

origin interpretation of most Ordovician chert deposits.

The quantitative results of this study show environmental and geographic patterns

in chert abundance, as well as temporal and spatial trends in chert distribution.

Considered in conjunction with various types of data of different nature developed by

previous studies for the Ordovician, the patterns and relationships revealed by this study

help to better understand the paleoecology of the Paleozoic silica secreting organisms,

their needs in terms silica and nutrients, and affinity for other physical and chemical

parameters, as well as their behavior regarding their response to changes in oceanic

circulation and climate parameters. They allow documentation and timing of an onshore-

offshore migration of silica secreting organisms in response to evolutionary patterns and

system perturbations.

The results also shed light onto the various factors that controlled the production,

accumulation and preservation of opal, such as organism productivity, sediment influx,

water temperature, salinity, host rock, or depositional environment. Interpreted in light of

90 actualistic and non-actualistic scenarios, the Ordovician data reveal functioning of a chert preservation model associated or not with volcanically-induced high organism productivity, within well aerated or anoxic bottom environments. Within this context, another important result is the documentation of chert accumulations on the shelf, a depositional environment not known to accumulate cherts in modern oceans.

91

Chapter 8

THE LOWER PALEOZOIC SILICA SECRETING ORGANISMS:

RADIOLARIANS AND SILICEOUS SPONGES

8.1. RADIOLARIANS

8.1.1. Systematics

Radiolarians are single-celled, siliceous-skeletonized planktonic marine ,

classified within the Superclass Actinopoda, Subphylum Sarcodina,

Sarcomastigophora, Kingdom Protista (De Wever et al., 2001). As fossil remains, only polycystine radiolarians (Order Polycystina) have been well preserved due to their opaline silica skeleton. Polycystines are represented by spumellar radiolarians (Suborder

Spumellaria) exhibiting a spherical symmetry, and nasellar radiolarians (Suborder

Nasellaria) characterized by bilateral symmetry.

Radiolaria studies greatly intensified with the initiation of the Drilling

Project in the 1960s, and a well-developed radiolarian systematics has emerged. Even though radiolarians have extant representatives, the systematics of the modern assemblages is based on skeleton morphology and not on cell anatomy and morphology as in classic , due to the difficulty in growing and studying radiolarians in the laboratory. Radiolarian biozonation became a powerful biostratigraphic tool not only for the Cenozoic and Mezozoic Eras (Aitchinson, 1998), but also for the upper Paleozoic

Eon. However, Jones and Murchy (1986) asserted that good taxonomic radiolarian

92 descriptions were achieved only for Cenozoic and upper Mesozoic faunas, and that 95% of the Paleozoic radiolarian assemblages were undescribed.

It is worth mentioning that the impossibility of extracting radiolarian skeletal remains from chert samples in view of a taxonomic study has represented a major impediment in the development of radiolarian taxonomy and biostratigraphy. Only after designing of a technique for extracting radiolarians from siliceous rocks, the hydrofluoric acid etching method (Pessagno and Newport, 1972; Dumitrica, 1970), did the biostratigraphy and taxonomy of radiolarian assemblages greatly improve.

8.1.2. Ecology

Modern radiolarians are among the marine zooplanktonic that are represented by various adapted to a wide ecological spectrum, inhabiting all of

Earth’s oceans from shallow water settings to deep waters of the abyssal plain. They range from 30 microns to 2 mm in diameter, and have various test shapes with attached spines. Radiolarians are filter feeders, predators, or live as symbionts with unicellular algae such as dinoflagellates

(http://www.ucmp.berkeley.edu/protista/radiolaria/rads.html). Radiolarians can be bacterivores, herbivores, symbiotrophes, detritivores, and etc (Casey, 1993). Although most radiolarians are solitary, some species are prone to form thread colonies up to a few meters in length (Casey 1993).

Modern radiolarians are most abundant in eutrophic, nutrient-rich warm surface waters principally of the equatorial zone, but are also present in subtropical and tropical environments. Their spatial distribution is controlled mainly by salinity, temperature, and

93

nutrient richness among other factors

(http://www.ucmp.berkeley.edu/protista/radiolaria/rads.html). However, as pointed out

by Abelman and Gowing (1997; p.4) “the knowledge of the vertical distribution of

polycistine radiolarian taxa in the water column is limited” so that the latitudinal and

basinal distribution patterns of radiolarians and a consistent relationship between global

distribution patterns and physico-chemical parameters of the seawater are lacking (Chang et al., 2003). Most controlled-environmental laboratory studies have analyzed the extent to which the seawater temperature, salinity and light intensity influence the growth, longevity, abundance, and opal productivity of modern radiolarian species (Anderson et al., 1989a,b,c; Anderson et al., 1990; Matsuoko and Anderson, 1992), and a few looked at the effect of dissolved silica levels (Sugiyama and Anderson, 1997). In this respect, studies of two modern radiolarian species demonstrate that polycystine radiolarians can be very effective in silica uptake, occurring abundantly in surface waters that exhibit low

concentrations of dissolved silica (ca. 1 µM), and yet mineralizing robust siliceous

skeletons (Sugiyama and Anderson, 1997). The four cultures of the study characterized

by different dissolved silica level (0 µM, 50 µM, 100 µM, and 150 µM) showed that: 1)

increased silica levels do not increase the mean weight of silicate deposited per shell (the

skeletal size), 2) nor the longevity, 3) the test morphology is not enhanced with

increasing dissolved silica, 4) silica levels as high as 150 µM strongly influence

negatively the longevity, growth, and weight gain of the two species of radiolarians,

probably having a poisoning effect that inhibits their metabolic activity and kills them.

However Sugiyama and Anderson (1997) pointed out that other environmental factors

94 such as temperature, salinity, and nutrients may represent more significant limiting factors on radiolarian growth and longevity than a low silica concentration level.

Anderson et al. (1989c) show that the optimum growth and longevity of Spongaster tetras tetras occurred at moderately warmer temperature and higher salinities (27.5ºC and

40‰ salinity); conversely, too high (above 33ºC) or too low a temperature (15ºC) has strong negative effect on longevity.

8.1.3. Evolution

Radiolarians go back in time as far as the Middle Cambrian (Knoll and Lipps,

1993; Won and Below, 1999) and represent the oldest planktonic microorganisms capable of secreting a mineral skeleton (Tolmacheva et el., 2001). Their ancestor is unknown.

The earliest-known reliably identified fossil radiolarians are known from Middle

Cambrian phosphorite-rich sediments that have accumulated in relatively shallow water facies of a siliciclastic platform in the Georgina basin, Australia (Won and Below, 1999).

Many radiolarian siliceous spicules were extracted from calcareous concretions of three different formations. Many other radiolarian-like fossils have been reported for the early and middle Cambrian times (Nazarov and Ormiston, 1986; Nazarov and Ormiston, 1993;

Rozanov and Zhuravlev, 1992; Iams and Stevens, 1988; White, 1986), but they have been regarded with skepticism by some authors (Aitchinson et al. 1998; Lipps, 1992).

The validity of the oldest reports of radiolarian skeletal remains is a debated problem because the siliceous skeletal components most frequently preserved within early Paleozoic facies are spicular skeleton elements, the morphology of which resembles

95

that of both siliceous sponge and radiolarian spicules. The hollow tube structure that

seems to be characteristic only for early Paleozoic radiolarians is a morphological style

common to sponge spicules and Silicoflagellata skeletons (Won and Below, 1999). As a

consequence, many Early and Middle Cambrian radiolarian-like spicules that have been

reported are regarded as or siliceous spicules.

The depositional environment where radiolarians emerged is a disputed subject

too – deep water versus shallow water. A few radiolarian-bearing strata from relatively

shallow water sedimentary facies of Middle and Upper Cambrian age, argue for a shallow

water radiation (Won and Below, 1999; Xiping et al., 1997). Xiping et al. (1997) in their

study of Late Cambrian Radiolaria of the Bitiao Formation (part of the Yangzte platform

margin in China), describe radiolarian species classified in the family Entactiniidae,

Suborder Spumellaria. The polycystine remains were found in shallow-water facies,

supporting the shallow water origin of the radiolarians. However, the depositional

environment of the oldest radiolarian fauna (Middle Cambrian of the Georgina Basin)

described by Won and Below (1999) is puzzling considering that during the Cambrian,

certain southern regions of the Georgina basin subsided at a much higher rate than the

rest, exhibiting rapid changes in facies (shales, limestones, dolomites and sandstones) and thickness (Brown et al., 1968). Other deep-sea radiolarian-bearing strata often associated

with remnants of oceanic crust in accretionary prisms of ancient subduction zones,

suggest a deep water (basinal) origin. According to Tolmacheva et al. (2001), the oldest

possible deep-sea radiolarian bearing strata could be those occurring in the basalt-

siliceous-terrigenous Zasuriya Formation of Upper Cambrian age, in the Altai Mountains

96

of central Asia. Aitchinson et al. (1998) describe lower Ordovician radiolarian-rich

sediments (spherical – spumellarian, and bi-polar forms) associated with ophiolitic

complexes, from the Little Port Complex in western Newfoundland (Tremadocian) and

from the Ballantrea Complex in Scotland (middle Arenigian). To date, the oldest well-

attested abyssal deep-sea biogenic chert was reported recently from south-central

Kazakhstan by Tolmacheva et al. (2001). The ribbon-banded radiolarian chert fomation is part of an ophiolitic sequence. The ages based on conodont biozonation showed that the sequence represents 15 m.y. of continous deep-sea biogenic siliceous sedimentation across the Cambrian-Ordovician boundary. In addition to radiolarians, megasclerites of hexactinellid sponges have been identified in the chert, illustrating the fact that siliceous sponges were already present in deep-sea environments by the Early Ordovician.

Another issue still open to controversy is the style of life that characterized the first radiolarians. The fact that the oldest radiolarian-like fossils have been found associated with and attached to sponge spicules (Bengston, 1986; Won and Below, 1999), along with laboratory cultures that show the ability of radiolarians to attach to surfaces, lend some support to the hypothesis that radiolarians originated from benthic forms

(Petrushevakaya, 1977; Rigby and Milsom, 2000). With regard to this, Nazarov and

Ormiston (1986) assert that Radiolaria stand for that group of organisms that might reflect the transition from a benthic to a planktic life habit. The fauna including definitive

Cambrian radiolarians is represented by speciemens of only one family, the Echidninidae, which are supposed to have lived as equatorial plankton (De Wever et al., 2001).

97

Regardless of the depositional environment in which radiolarians originated in the

Middle Cambrian, by the Lower Ordovician radiolarians inhabited shelf (Blome et al.,

1995a; Renz, 1990; Kozur et al., 1996), as well as deep-water settings (Aitchinson et al.,

1998; Tolmacheva et al., 2001). However, the present study shows that most Lower

Ordovician radiolarian remains are predominantly associated with bedded cherts of deep

water sedimentary sequences, a style that probably represents a continuation from the

Upper Cambrian, and possibly the Middle Cambrian. Twenty million years later, all

Middle Ordovician radiolarian-bearing formations are exclusively bedded, chert-

dominated deep water facies, suggesting a shelf to slope/basin migration from the Lower

to Middle Ordovician. This status is preserved in the Upper Ordovician and apparently in

the Silurian too (Tomescu et al., 2003).

During the Ordovician, radiolarians became the dominant component of the

microzooplanktonic fauna (Tappan and Loeblich, 1973). Radiolarian diversity increased

during the Ordovician when several groups were present, dominated by spherical porous

polycystines (Nazarov and Ormiston, 1986; De Wever et al., 1994). The Ordovician

radiolarians were diverse and widespread enough to allow interbasinal biostratigraphic

correlations between Eurasia and North America (Nazarov and Ormiston 1993), and a

radiolarian biozonation (Noble and Aitchinson, 2000). Nonetheless, the Ordovician-

Silurian radiolarian biota records a slow evolution of spherical Spumellarians since their emergence, compared with Middle and Late Paleozoic biotas (Nazarov and Ormiston,

1986; Vishnevskaya, 1997). If the Ordovician and Silurian radiolarian assemblages are characteristic at the series level, during the Middle and Upper Paleozoic they developed a

98

higher diversity, allowing biostratigraphic subdivision at the stage level, and even at a

finer scale for assemblages (Nazarov and Ormiston, 1988; Nazarov and

Ormiston 1993). Vishnevskaya (1997) studying Paleozoic and Mesozoic radiolarian

assemblages of the Northwestern Pacific Rim (including eastern and northeastern

Russia), distinguished four Paleozoic megacycles in the development of radiolarian fauna, of which the first, represented by the Ordovician-Silurian interval, is characterized

by a low diversity compared to the other Upper Paleozoic megacycles. Vishnevskaya’s

worldwide compilation table (Fig. 11.1) based on radiolarian literature published until

1995, points out that world-wide species diversity of radiolarians was much higher during the Ordovician (10 species in the Tremadocian and up to 100 species in the Caradocian) than during the previous and subsequent periods - Cambrian with around 20 species, and

Silurian with approximately 30-40 species. According to her data, the Middle Ordovician

species diversity is surpassed only by Late (up to 125 species) and Lower

Carboniferous (around 115 species) faunas in the Paleozoic.

Contrary to the view of a low Ordovician diversity, Renz (1990), analyzing the

Ordovician radiolarian assemblages from Nevada and Newfoundland at the family level,

reported a significant radiolarian diversification as a component part of Ordovician

evolutionary changes. His study showed that by the Upper Ordovician, the dominance of

spherical polycystines (family Inaniguttidae of Nazarov and Ormiston, 1986) is replaced

by a group characterized by bilateral symmetry and similar to family Palaeoscenidiidae, a

family considered by Nazarov and Ormiston (1986) and Jones and Murchey (1986) to

have emerged in the Middle Silurian. The Upper Ordovician radiolarian assemblages

99 inhabiting Gondwana (Australia) and Laurentia (Nevada, Newfoundland), but situated in the same paleolatitudinal range (20º from the equator), have distinct characteristics

(Renz, 1990), suggesting at least a continental-level endemism.

Whether or not the Late Ordovician glaciation affected radiolarian faunas is controversial. In contrast to the lack of major radiolarian suggested by Noble et al. (1997), Vishnevskaya (1997) remarks a noteworthy radiolarian extinction across the

Ordovician-Silurian boundary (Ashgillian – Llandoverian), the first one among the five identified Paleozoic extinctions. Across this boundary the radiolarian diversity drops to

20 species. Following this drop, the rest of the Silurian witnesses a slight increase in diversity upwards of 60 species (Vishnevskaya, 1997; Vishnevskaya Kostyuchenko and

2000).

8.1.4. Geologic importance

Radiolarians exemplify the oldest and the only skeletonized planktonic organisms preserved in sedimentary sequences of all Phanerozoic systems, inhabiting the oceanic realm since at least the middle Cambrian and until Recent times. Polycystine radiolarians also represent the only skeletonized that have been found in Cambrian rocks, albeit in low abundance (Horodyski et al., 1992). Numerous Lower Paleozoic radiolarian assemblages are incorporated into the accretionary prisms associated with subduction zones (Aitchinson, 1998). Because Paleozoic radiolarian-bearing chert deposits are commonly associated with ophiolitic complexes, many authors emphasize the use of radiolarians as a biostratigraphic tool in dating lower Paleozoic orogenic belts devoid of other biostratigraphic potential and consequently lacking a good age control (Grunau,

100

1965, Aitchinson et al., 1998; Blome et al. 1995a,b; Nazarov and Ormiston 1993).

According to Noble and Aitchinson (2000), early Paleozoic radiolarian faunas include several test morphotypes. A Lower Paleozoic radiolarian biostratigraphy and biozonation was set up by Nazarov and Ormiston (1993) for Eurasia, and extended subsequently toward the North American continent, most recently by Noble and Aitchinson (2000).

The high sensitivity of radiolarians to water temperature and chemistry has

rendered them an excellent paleoceanographic and paleoclimatologic marker, mainly for

the Cenozoic era. Distribution patterns and abundance of distinct modern radiolarian

assemblages are tightly linked to the amount of nutrients dissolved in the seawater and to

water depth (Abelmann and Gowing, 1997; Kruglikova, 1993; Chang et al., 2003), in

addition to warm temperature (around 27ºC) and salinity (35-40‰) (Anderson et

al.,1989a,b,c; 1990; Matsuoko and Anderson, 1992; Sugiyama and Anderson, 1997).

Based on an actualistic scenario, the abundance of radiolarian-bearing strata has been

used as an indicator of paleo-upwelling and high productivity areas for the Phanerozoic

(Hein and Parrish, 1987). Also, radiolarian-bearing chert deposits preserved since the

Ordovician have been used successfully as an indicator of deep-water settings (distal basinal or oceanic environments associated with orogenic belts), as well as of cratonic seas of moderate depths (Aubouin, 1965; Ormiston, 1993). Vishnevskaya (1997) has shown that evolutionary changes undergone by radiolarians throughout the Phanerozoic can be linked to sea level changes, paleoceanographic regime, and plate tectonics events.

Modern Spumellarian species inhabit mainly shallow waters, as opposed to

Nasellarian species that inhabit deep waters. This ecological separation has been used to

101

deduce nearshore versus offshore environments and to propose sea level fluctuations by

calculating the Spumellarian/Nasellarian ratio (Casey, 1993). However, this is possible

only for sediments younger than the because the oldest unquestioned

Nasselarians date from the Middle Jurassic (Vishnevskaya, 1997).

8.2. SILICEOUS SPONGES

8.2.1. Systematics

Sponges, the simplest metazoan organisms (Wainright et al. 1993; Conway,

1994), feature specialized cells, but these are not grouped into distinct tissues. Sponges

are grouped into Phylum Porifera, Kingdom Animalia. The sponge classification system

is based, as in the case of radiolarians, mainly on skeleton morphology. Presently,

phylum Porifera comprises 4 classes: Demospongea, Hexactinellida, Calcarea, and

Sclerospongea (Rigby 1983). Fossil siliceous sponges have siliceous spicules as

structural skeletal elements and are represented by Class Hexactinellidae or

Hyalospongea (glass sponges), and Class Demospongea.

8.2.2. Ecology

Sponges are sessile, benthic filter-feeding metazoans. Presently, Demosponges represent

approximately 95% of all sponges (Rigby 1983), with over 600 living genera that inhabit almost any aquatic environment from shallow to deep settings (Prothero, 2004). In the past, including the Ordovician and the Silurian, Demosponges were important reef- building organisms or major components of the reef communities (Finks, 1970; Rigby,

1983). Even though most of the modern species of sponges prefer clear, shallow marine

102

waters, siliceous sponges prefer deep quiet waters up to 5000 m deep (Prothero, 2004). If we consider the hexactinellids versus lithistid demosponges, there is a differentiation in their most favorable living habitat. Most living Demosponges prefer relatively warm, shallow, tropical seas. However, they are known to inhabit environmental settings as

deep as 1500 m (Rigby, 1983; van Soest, 1994), and some modern Demospongea species

(e.g., Halichondria panicea) live in shallow-waters of high latitude environments (55˚ N)

as cold as 10˚C, such as those offered by the Baltic Sea (Reincke and Barthel, 1997). This

siliceous sponge genus with a cosmopolitan distribution (van Soest, 1994) represents an

important epifaunal component of the red algae zone at depths ranging from 4 to 11 m.

As such, it could compete with diatoms for the available dissolved silica in the seawater,

but this is a topic still open to controversy. On the other hand, Hexactinellids inhabit

mainly the upper bathyal zone (between 200 and 2000 m in depth, Noble, 1997;

Tabachnick, 1994), and are found also toward the deepest habitats such as abyssal and

hadal niches (below 6000 m) (Hartman 1983; Tabachnick, 1994).

The differentiation between Demosponges and Hexactinellids in terms of their

most favorable living habitat is seen as a result of their different requirements for

dissolved silica levels in the seawater: hexactinellid sponges need higher Si(OH)4 levels than demosponges. Consequently, the deep water habitats of modern hexactinellids are known to be the silica-rich environments (Tréguer et al., 1995; Nelson et al., 1995), or shallow water settings that are rich in dissolved silica, such as those of higher latitudes, or microhabitats with enhanced Si supply via rivers (Maldonado et al., 1999; Boury-Esnalut and Vacelet, 1994).

103

Siliceous sponges generally need higher dissolved silica levels than the diatoms and radiolarians and consequently the silica limitation characteristic for neritic environments could explain the Cenozoic migration of most siliceous sponges toward deeper silica-rich settings (Maldonado et al., 1999). Under very low silica concentration environments, as low as 0.741 ± 0.133 µM (the mean ± s.e. of a silica concentration range

from 0.03 to 4.5 µM), siliceous sponges are capable of secreting only very small and thin

needle-like spicules (Maldonado et al., 1999). Experiments demonstrated that there is an

asymptotic relationship between the silica uptake rates in Demospongea and the amount

of dissolved silica in the seawater (Reincke and Barthel, 1997; Frøhlich and Barthel,

1997; Maldonado et al., 1999) – the silica uptake rate increases asymptotically with

increasing silica concentration. The higher the available dissolved silica level, the more

complex and varied the spicules that are secreted by siliceous sponges (Maldonado et al.,

1999). In contrast to diatoms, the saturation level of the Demosponge genus Halichondria

panicea is much higher (above 150 µM Si-concentration) and silica uptake rates are much lower (17 µmol h-1g-1 ash-free dry weight) (Frøhlich and Barthel, 1997). By comparison, the saturation point of some diatom species is reached at 10µM, with an

uptake rate of 0.001 µmol Si h-1cell-1.

Is silica or nutrient limitation of greater importance? This a question raised by

many authors trying to explain the growth and evolution of siliceous sponges. However, modern studies still argue about which parameter dominates. The limitation

growth theory is supported experimentally by the work of Reincke and Barthel (1997)

and Maldonado et al. (1999). However, Maldonado et al. (1999) studying cultures of

104

Crambe crambe, a demosponge species characteristic of relatively low latitude coastal waters in tropical/subtropical enviroments (e.g., the Mediterranean Sea, between approximately 30˚ and 45˚ north), notice that overly high seawater silica concentrations

(100 µM) may inhibit silica uptake. This observation does not fit the results obtained by

Reincke and Barthel (1997) in their experiments, probably because they used a different demosponge taxon also characteristic for shallow waters, but populating much higher geographical latitudes, respectively around 55˚ north.

The nutrition limitation growth theory is supported by experiments of Frøhlich and Barthel (1997). They showed experimentally that silica uptake is an energy- consuming process and therefore in a nutrient-starved environment the biogenic secretion of dissolved silica decreased by 15%. However, their experiment does not show what happens to siliceous sponges if the low nutrient level goes below a certain threshold (do siliceous sponges cease silica uptake and die?).

Normally, at higher temperatures higher silica uptake rates should occur. With increasing temperature the silica cycling is increased by higher dissolution rates of the skeletons that generate higher dissolved silica levels, and therefore trigger higher silica uptake rates. However, it was shown that an increase from 10˚C to 15˚C, did not produce a measurable increase in silica uptake rate for modern demosponges (Frøhlich and

Barthel, 1997). This could be explained by differences in the magnitude of the effects of temperature changes on the rates of silica dissolution and silica uptake. A higher increase in temperature would perhaps trigger perceivably higher silica uptake rates.

105

The role of the siliceous sponges in the modern silica cycle is viewed as

insignificant because of the silica cycle domination by diatoms. Maliva et al. (1989) claim that siliceous sponges once exerted the main control on the Si cycle in neritic environments, and that throughout the Phanerozoic siliceous sponges have undergone changes in their environment.

8.2.3. Evolution

Phylogenetic theories propose that metazoans evolved from a by acquiring specialized multicellularity (Cowen, 2000), but their phylogeny is still an issue open to controversy. Molecular work based on comparisons of small subunit ribosomal RNA

(rRNA) sequences (Wainright et al. 1993) shows that within the Metazoa sponges were the first to radiate. Different molecular studies of sponge phylogeny since the late 1980s support this interpretation and suggest that sponges have paraphyletic rather than monophyletic origins (Borchiellini et al., 2000; Borchiellini et al., 2001).

All the major classes of sponges (Calcarea, Hexactinellida and Demospongea) are known from at least the Cambrian. Porifera has been considered a conservative group of organisms because sponge taxa display long geologic ranges (Carrera and Rigby, 1999).

Siliceous sponges occur as far back as the , as a part of the or

Vendian faunas. According to Wood et al. (2002), calcareous sponges could be as old as

549 Ma (Neoproterozoic). They describe a calcite-biomineralized large, modular metazoan from a shallow marine carbonate - siliciclastic sequence (the ) in southern . The fossil, dated at 548.8 ± 8 Ma B.P. (Grotzinger et al. 1995), presents some morphological structures of both poriferan and cnidarian affinity.

106

The oldest well-preserved siliceous axial spicules discovered to date were

identified as monaxonid Demosponges, and dated at approximately of 580 Ma (Li et al.

1998), a time of expansion of the Ediacaran fauna. They are preserved within the

Vendian Doushantuo phosphatic sequence in south China, central Guizhou. The shape of

the siliceous spicules is mostly globular, but a few of them are tubular. The identification

of the oldest well-documented siliceous sponge spicules as Demosponges suggests that a

demosponge was the ancestor of the sponges.

Younger Vendian spicules that are unequivocally from sponges, classified as hexactinellids and not as demosponges, have been described by Brasier et al. (1997).

These spicules exhibit various morphologies demonstrating that skeletal diversity had been achieved by Hexactinellids by the latest . They were discovered in

Mongolia within the upper part of the Tsagaan Formation in a limestone unit overlying a phoshorite-chert-black shale succession. The unit was dated as late Ediacaran age, 543–

549 Ma (Grotzinger et al., 1995). This finding refuted the earlier theory according to which hexactinellids appeared during the early Cambrian (Finks, 1970; Finks, 1983) or

Middle Cambrian (Webby, 1984a). Ediacaran-age siliceous sponge spicules attributed to

Hexactinellids, have also been reported from South Australia, in the Pound Subgroup

(Gehling and Rigby, 1996), but their origin is disputed (Li et al., 1998).

The issue of the ancestry of either Demospongea or Hexactinellida within Porifera is still debated. However, the emergence of siliceous preceding the calcareous biomineralization in sponges is widely accepted (Li et al., 1998; Brasier et al.,

1997).

107

Based on associated lithologies, it was shown that the Late siliceous

sponges inhabited eutrophic environments within shallow seas, in contrast to early

Cambrian calcareous sponges that inhabited “more interior” oligotrophic settings (Li et

al., 1998; Brasier, 1992). Speculations are made on a eutrophic outer shelf origin of the

Neoproterozic mineral skeleton faunas, closely related to high Si, P and Ca

concentrations in the water column, and on a shoreward migration of these faunas during

the Cambrian transgression (Brasier, 1992; Brasier et al., 1997). The shallow water

radiation of sponges is supported by the widely accepted theory of the shallow water origin of life (Crimes, 1992). In contrast, Webby (1984a) proposes that sponges emerged in deeper slope environments during Late Precambrian – early Cambrian times, and migrated and dispersed subsequently onto carbonate platform habitats “after the demise of archaeocyaths”. Since the Neoproterozoic sponge faunas have been characterized by high tiering levels (Yuan et al. 2002), demonstrating their evolved ability in extracting nutrients from the water column. By the Neoproterozoic, shallow waters were dominated by associations and subordinate siliceous sponges, in contrast to deep waters characterized by basin siliceous assemblages (Finks, 1960).

In the Early Cambrian, both monaxonic Demosponges and Hexactinellids already represented a relatively diversified component of the benthic fauna (Finks, 1970), inhabiting both shallow and deep-water settings. Numerous siliceous sponge fossils have been discovered in China (early Cambrian Hexactinellids and Lithistids; Zhang and Pratt,

1994), South Australia (early Late Cambrian hexactinellid type spicules ; Bengston,

1986), and the Siberian platform (Early Cambrian Hexactinellids and Demosponges;

108

Rozanov and Zhuravlev, 1992). In the Lower Cambrian Chengjian Formation of China

15 genera and 30 species of Demosponges were identified by Li et al. (1998). Yuan et al.

(2002) in their study of early Cambrian epifaunal tiers from China, identified giant hexactinellids and smaller demosponges. Zhang and Pratt (1994), studying Early

Cambrian siliceous spicules (Hexactinellidae and Lithistidae) from argillaceous

limestones interpreted as a slope deposit in Shaanxi (south China), asserted that siliceous

sponges became moderately widespread in low-energy, offshore environments in China,

Australia, and Siberia by the Atdabanian (Lower Cambrian). Their theory supports the

proposal that siliceous sponges emerged in deep water setting (Webby, 1984a). Won and

Below (1999) mention carbonate concretions rich in siliceous sponge spicules from

Middle Cambrian sediments that accumulated in the relatively shallow water facies of a

siliciclastic platform in the Georgina basin, Australia. Kruse (1983) identified siliceous

sponges from the Georgina basin as lithistid demosponge forms. Deep water

Hexactinellid siliceous sponges are known from a Late Cambrian biogenic chert deposit,

part of an ophiolitic sequence, in Kazakhstan (Tolmacheva et al. 2001).

During the Ordovician the siliceous sponge faunas, mainly Lithistid demosponges

(Finks, 1970; Rigby 1983), became even more complex, diversified, and widespread

being represented not only by endemic but also by cosmopolitan genera. According to

Finks (1970), by the Middle Ordovician lithistids reached a relatively high diversity (17

genera), representing an important reef-building component. Overall, Ordovician sponge records point to a higher diversity of demosponges compared to hexactinellids (Carrera and Rigby, 1999). By the end of the Ordovician, several orders of sponges with multiple

109

suborders had already evolved, demonstrating a high degree of diversification. According

to Carrera and Rigby (1999), the most abundant Ordovician endemic sponge fauna occurs

in Australia, especially in the New South Wales area, with endemism levels being lower

in the Laurentian Appalachians and China. Calcareous sponges, especially sphinctozoan

sponges, almost equaled the siliceous sponge diversity in the Middle and Upper

Ordovician.

The paleogeographic distribution patterns of Ordovician sponges illustrates a

change in sponge faunas during the Ordovician period (Carrera and Rigby, 1999). Early

Ordovician sponge faunas have relatively low diversity and are dominated by

demosponges. Middle Ordovician faunas, also dominated by demosponges, display a

higher diversity characterized by two main associations – Appalachian faunas (including

South China and the Argentine Precordillera), and Great Basin faunas. The Late

Ordovician exhibits pronounced increases in diversity of both siliceous and calcareous sponges and provincialism with two biogeographic associations: the Pacific association

(western North America and New South Wales of Australia), and the Atlantic association

(Midcontinent Laurentia and Baltica).

Rigby and Webby (1988) analyzed Late Ordovician sponge associations from the

Malongulli Formation of Central New South Wales, Australia, and described 34 genera

divided into 44 species of siliceous sponges (both demosponges and hexcatinellids, with

the lithistid type dominating) and calcareous sponges. According to the authors, this sponge assemblage exemplifies a unique record of deeper-water carbonate environments bordered by an island-arc shelf sequence. This Late Ordovician New South Wales sponge

110

association also suggests a clear-cut differentiation between calcareous sponges that

dominated shallow waters, and siliceous sponges (both demosponges and hexactinellids) that dominated slope-basin environments (Rigby and Webby, 1988).

Within a shallowing interval of the lower Hanson Creek Formation (late Upper

Ordovician) Finney et al. (1999) recognized a faunal turnover in siliceous sponges. The change from deep water, hexactinellid-dominated assemblages, to shallow, subtidal lithistid-dominated assemblages provides an example of the association between sponge types and depositional environment and illustrates the responsiveness of sponges to environmental changes.

111

Chapter 9

DISCUSSION: THE ORIGIN OF THE ORDOVICIAN CHERTS - ORGANIC VS.

INORGANIC GENESIS

9.1. Previous theories on chert origin

The inorganic formation of chert through direct precipitation from seawater

supersaturated with respect to silica seems the most appropriate mechanism to explain the

formation of Precambrian cherts. Conversely, the genesis of most chert deposits from the mid-Paleozoic on is viewed and widely accepted as a biologically induced process with the common inference on association of bedded cherts with radiolarians (Grunau, 1965;

Ramsay, 1973; Hein and Parrish, 1987), and nodular cherts with siliceous sponges

(Maliva and Siever, 1989; Maliva et al., 1989) are genetic associations. However, controversy still exists regarding the dominant mechanism responsible for the formation of chert deposits during the Early Paleozoic, principally the Cambrian and to a lesser extent the Ordovician.

The Proterozoic cherts are unanimously thought to be inorganically generated through the precipitation of silica from a silica supersaturated sea, based on dearth of skeletal structures within cherts owing to the absence of major silica secreting biotas that could balance the excess of silicic acid in the seawater (Siever, 1957). Several workers have proposed that the silica cycle has changed through time, from an ocean supersatured with respect to silica during the Precambrian, toward a silica-undersaturated ocean during 112

the Phanerozoic, as siliceous organisms evolved (Siever, 1957, 1962; 1992; Maliva et al.,

1989).

The organic theory proposed as the main mechanism responsible for the formation of most chert deposits, principally of bedded cherts, from the mid-Paleozoic on, is widely accepted because the Paleozoic silica secreting organisms had reached an evolutionary level that could explain their ability to keep the silica concentrations undersaturated at all times. This conclusion is based on the high biological diversity

(Nazarov and Ormiston, 1986; Vishnevskaya, 1997), the abundance of skeletal remains

preserved in cherts during the Late Silurian, Devonian, and Permian, and

the inference that dissolved silica concentrations might have been highest where

sediments contained more biogenic silica. The latter is based on a modern scenario

(Calvert, 1983). The lack of any consistent inorganic process of sufficiently high

magnitude to account for the genesis of areally extensive cherts and the lack of any

coherent association patterns between sedimentary sequences supposed to be responsible for high silica inputs and chert deposits further support a biogenic origin of Mid- to Late

Paleozoic chert.

The inorganic genesis as a dominant mechanism responsible for the formation of

Early Paleozoic cherts is supported mainly by a higher magnitude of the volcanic activity

(Ronov et al., 1980) and of silicate rock weathering (François et al., 1993). This theory is

consistent with the presence in certain bedded cherts of minerals regarded as volcanic

alteration products, such as montmorillonite, palygorskite, sepiolite, and clinoptilolite

(Wise and Weaver, 1974) and the hypothesis that there is a direct relationship between 113

the abundance of clay minerals in sediments and silica concentrations (Lancelot, 1973a,b;

La Porta, 1998a). The association of certain siliceous deposits such as the bedded cherts

with volcanogenic massive sulfide (VMS), (Grenne and Slack, 2003) and the dearth or low preservation of skeletal remains in many cherts (Von Rad and Rösch, 1974),

principally for the Cambrian, also suggests an inorganic mechanism for chert formation.

This is consistent with the calculation that seawater concentration with respect to

dissolved silica was high enough to allow the precipitation of opal-CT as chert nodules

(Siever, 1991). The Early Paleozoic oceans had high endogenous and exogenous silica

input levels driven by high continental weathering rates during the Cambrian (as inferred

from the isotopic strontium curve; Qing et al., 1998) and energetic marine volcanic

activity mainly during the Ordovician. The general assumption is that silica-secreting

organisms during the Cambrian and to a lesser extent during the Ordovician, were at the

beginning of their development, and had not reached the level of evolution and

abundance beyond which they could regulate the marine silica cycle and control the

biogenic burial of silica.

Siliceous sponges originated in the Neoproterozoic but were probably not able to

influence the marine silica cycle by lowering the silica concentration under the

precipitation level and produce biogenic cherts at a global scale until at least the Middle

Cambrian, when another group of organisms capable of extracting silica from the

seawater, the radiolarians, originated (Won and Below, 1999). The Cambrian is

moderately rich in bedded chert deposits, most of them formed in shelf environments

(Hein and Parrish, 1987; Kidder and Erwin, 2001). There is no survey regarding the 114

faunal content of these cherts and the question remains: Did radiolarians and siliceous

sponges induce their formation?

The earliest reliably identified fossil occurrence of radiolarians is known from the

Middle Cambrian (Won and Below, 1999), but hypotheses were put forward for their

emergence earlier, in the Early Cambrian (Ormiston, 1993). Regarding the depositional

environment where radiolarians originated, evidence of Middle Cambrian radiolarian- bearing sedimentary sequences argues for a shallow water radiation (Won and Below,

1999), a hypothesis supported by Late Cambrian radiolarian-bearing facies (Xiping et al.,

1997), and also in harmony with the shallow water origination of most major groups of

organisms (Webby, 1984a; Sepkoski and Miller, 1985). On the other hand, the deep-sea

origination hypothesis is supported by a Late Cambrian basalt-siliceous-terrigenous

formation bearing well identified radiolarians (Tolmacheva et al., 2001), the skepticism

regarding the identification of the earlier radiolarian-like fossils for the Early and Middle

Cambrian times (Aitchinson et al., 1998), and the consideration that during the Cambrian

certain southern regions of the Georgina basin, which is the area of the earliest-known

reliably identified radiolarians (Won and Below, 1999), subsided at a much higher rate

than the rest of the basin, exhibiting rapid changes in facies (shales, limestones, dolomites

and sandstones) and thickness (Brown et al., 1968), and thus casting doubt over the

hypothesized shallow environment.

The shallow water origination of radiolarians during the Cambrian would be

consistent with the more abundant shelf bedded cherts during the same period (Kidder

and Mumma, 2003), only if the association of bedded cherts and radiolarians was an 115

interconnected genetic process as thought by many authors, and demonstrated herein at

least from the Ordovician on. It could also be that siliceous sponges were chiefly

responsible for chert formation in both the Early and Middle Cambrian, and that some

radiolarians accumulated in sponge cherts. No taxonomic evidence of radiolarian fossils

exists for the Lower Cambrian, when bedded cherts were forming as well (Hein and

Parrish, 1987). Yet, by the Early Cambrian, both monaxonic Demosponges and

Hexactinellids already represented a relatively diversified component of the benthic fauna, inhabiting both shallow and deep-water environments (Finks, 1970). Thus it is possible that by the Early Cambrian the siliceous sponge fauna might had came to the point where it could reduce silica concentrations under the saturation levels, driving the formation of the first biogenic cherts of both types – bedded and nodular. It is very likely that the Cambrian witnessed a shift from a dominantly inorganic silica precipitation to a globally biogenic induced chert formation, and from a siliceous sponge dominated silica burial to a radiolarian-sponge dominated silica deposition, so that during the Ordovician

the silica burial was mainly biogenically controlled. The Early-Middle Cambrian shift

from a dominated inorganic silica precipitation to a biologically induced process is

supported by the deleterious effect of high silica concentration levels on the behavior of

siliceous sponges and radiolarians, finding discovered through experiments on modern

faunas (Sugyiama and Anderson, 1997; Maldonado et al., 1999). Some Cambrian nodular

chert bearing formations have been suggested of inorganic origin based on lithological

associations such as the presence of clay-rich seams at chert contacts and of clay 116 inclusions within the silica matrix (La Porta, 1998a,b), but these deposits are not well studied.

9.2. Evidence from this study

The data that support the organic origin of Paleozoic chert deposits, and the association patterns between the type of skeletal remains (radiolarians vs. siliceous sponges), the chert style (bedded vs. nodular), and the depostional environment in which they form (peritidal/lagoonal vs. shelf vs. slope/basin) is mainly quanlitative. The few compilations of worldwide Phanerozoic bedded cherts (Grunau, 1965; Ramsay, 1973,

Hein and Parrish, 1987) show the affinity and a possibly genetic relationship between radiolarians and bedded cherts. As regards the possible association between nodular cherts and siliceous sponges, to date there is no systematic survey, for any geological period, of the nodular cherts and their corresponding content in biogenic siliceous remains. The present study brings the first quantitative evidence in support of the biogenic origin of cherts and genetically-dependent association patterns.

The organic origin of most Ordovician cherts are organic in origin is supported by two aspects of this study. First, approximately 38% of the Ordovician chert formations contain siliceous skeletal remains, mainly radiolarians and subsequently siliceous sponges. Second, the onshore-offshore migration of the silica secreting organisms perceived through the rock record of the siliceous organism bearing chert formations, is accompanied by an onshore-offshore change in the locus of silica burial, perceived in the stratigraphic record of cherts that lack preserved siliceous skeletal remains. The 117

systematic analysis of chert content in terms of biogenic remains demonstrates that Early

Paleozoic silica secreting organisms have been underestimated with respect to their role

within the marine silica cycle and silica burial. It is worth mentioning that the

impossibility of extracting radiolarian and sponge skeletal remains from chert samples

has undoubtedly represented an impediment in the development of radiolarian and

siliceous sponge taxonomy and biostratigraphy, and hence in discovering which cherts

bear fragments of silica secreting organisms, until the 1970s. Many of the Ordovician

chert deposits were identified and described prior to the discovery of a technique capable

of extracting siliceous organisms from siliceous rocks, the hydrofluoric acid etching

method (Pessagno and Newport, 1972; Dumitrica, 1970). Based on this, it seems

probable that the number of Ordovician cherts bearing siliceous skeletal remains might

actually be higher than found by this study.

The various relationships examined in this study (Chapter 7) reveals four

statistically significant associations: first, a strong affinity between radiolarians and deep-

water (slope/basinal) bedded cherts; second, siliceous sponge spicules can be found in

either bedded or nodular cherts formed preferentially in shallow environments of

peritidal/lagoonal and shelf settings; third, bedded cherts form preferentially in

slope/basinal environments, and subordinately in shallower settings; and fourth, nodular

cherts form preferentially in shallow waters of shelf and peritidal/lagoonal environments.

Even though this study reveals no single example of bedded chert deposits containing only sponge spicules, the statistical analysis reveals an equal tendency of siliceous sponges occuring in either type of chert deposit but mainly with shallow water 118 deposits (peritidal/lagoonal and shelf). This finding supports the common conception that bedded cherts formed principally in slope/basinal environments within a siliciclastic regime represent primary accumulations of siliceous tests of silica secreting planktonic organisms (Grunau, 1965; Aubouin, 1965; Hein and Parrish, 1987), and that nodular cherts represent diagenetic products whose silica source is represented principally by siliceous sponges (Siever, 1962; Maliva and Siever, 1989). It also emphasizes that the formation of some shelf bedded cherts might by formed by siliceous sponges. None of the chert occurrences from this study identified as bedded cherts that formed on the shelf are associated with any volcanic elements or other potentially high silica generating factors that could explain their inorganic formation. However, for a couple of Lower

Ordovician bedded cherts associated with peritidal/lagoonal environments, the biogenic origin seems less probable (Pettijohn, 1975).

The present study shows that chert nodule formation occurred in the shallow water settings of the shelf and peritidal/lagoonal environments, and not in slope/basinal environments. This suggests there was a genetic relationship between siliceous sponges and nodular cherts. The question is: why is the formation of nodular cherts mainly confined to these shallow settings largely populated by siliceous sponges, when the required inorganic precipitation level of opal-CT, whcih is much lower than that of biogenic opal (40-60 mg/L SiO2 at 25°C) and which was achieved during the Ordovician

(Siever, 1991), indicates that the deeper, much richer silica environments would be more suited for their inorganic chert formation regardless of the presence of benthic silica secreting organisms? Moreover, the fact that many shallow water cherts have been found 119 containing siliceous sponge spicules, in contrast with the scarcity of sponge spicules in deep chert deposits, indicates an intrinsic relationship between the formation of nodular cherts and siliceous sponges inhabiting the proximal shallow environments.

The transition from the Lower to the Middle Ordovician witnesses a migration of the formations bearing biogenic siliceous remains toward deeper environments, concomitant with a sharp decline in the peritidal/lagoonal and shelf cherty facies. Why would the general pattern of the marine silica burial follow the migration of the silica secreting organisms? The conjunction of the two events suggests that there is a direct relationship between chert formation and the silica secreting organisms inhabiting the most proximal environments. One could argue that during the early Middle Ordovician the sharp sea level drop (Finney, 1997) exposed the shelf area, so that the change in the locus of silica burial could be an artifact of a sea level lowering, and not a consequence of the migration of the silica secreting organisms from the peritidal/lagoonal and shelf environments. An argument against this view would be given by the total retreat of radiolarians from shelf to basinal settings (indicated by the absence of Middle Ordovician shelf radiolarian-bearing cherts) and the sharp drop in the abundance of the silicified oolites (Table 7.7) characteristic of peritidal/lagoonal settings. The exposure of an important part of the shelf area, as the one that has occurred at the Lower-Middle

Ordovician boundary, would diminish the area extent of the optimal environment for the formation of oolites, and consequently fewer silicified oolites would be generated. Did this surface reduction generate the drop in silicified oolites, while siliceous sponges were still populating such very shallow environments? An answer to this question is provided 120

by the depositional distribution of the chert facies in the Caradocian. During the

Caradocian, when the Ordovician sea level was at its highest (Ross and Ross, 1992), the

locus of silica burial is mainly confined to shelf and basinal settings, whereas the

peritidal/lagoonal environments are characterized by just two silicified oolite deposits.

It is true that the presence of skeletal remains does not guarantee the organic

genesis of a deposit, but neither does the lack of such remains demonstrate inorganic

precipitation of silica. While the preservation of biogenic siliceous fragments in a chert

can introduce a bias toward the idea that the chert might be of biogenic origin, the lack of any skeletal structures gives equal chances to both organic and inorganic origins, as valid mechanisms. Multiple types of evidence are needed in order to decide between these two origins, and here I present the lines of evidence in support of biogenic origin for most chert deposits:

1) The siliceous skeletal remains are found associated with cherty rocks only.

2) The onshore-offshore silica secreting organism migration is accompanied by a corresponding change in the locus of silica burial, or in other words the depositional patterns in which cherts occur are in concert with evolutionary events, such as onshore- offshore faunal migrations.

3) The Early Paleozoic, silica-secreting organisms represented by radiolarians and siliceous sponges were relatively well diversified, abundant, and globally dispersed with efficient capabilities of silica uptake (as demonstrated by modern studies), with a high ability to balance globally the excess of silicic acid in seawater. Additionally, studies of modern species of radiolarians and siliceous sponges show that while they have the 121

capacity to extract dissolved silica from the seawater at very low concentrations (ca.

1µM), high dissolved silica levels in the seawater (100 - 150 µM) might be toxic

(Sugiyama and Anderson, 1997; Maldonado et al., 1999) (Chapter 8). Under these circumstances, it is im probable that the dissolved silica concentration reached saturated or supersaturated levels at the basin level, favoring in this way the inorganic formation of areally extensive cherts. Even though radiolarians and siliceous sponges are not as efficient as diatoms in processing silica uptake, they still secrete silica and serve as an effective “micro site” of silica bio-precipitation. Through the aggregation of their skeletal parts silica-rich sites develop that, in turn, act as a silica source much higher than the amount of the dissolved silica from the seawater. Moreover, thermodynamically it is easier to export enough silica to generate the formation of a chert deposit via siliceous organisms that are so efficient in silica uptake at the very low concentrations, than to raise the dissolved silica level in the seawater at concentrations so high that a direct precipitation of a chert deposit could occur.

4) Limestone is a very common rock in the Phanerozoic record. However only some of them contain cherts even though, there is a great affinity between carbonate rocks and certification (Kastner et al., 1977) (Chapter 5). The selective occurrence of chert in only some limestones supports the biogenic origin as does the loss of chert nodules from most Cenozoic limestones when sponges moved to deeper waters.

5) The spatial distribution of chert deposits does not not correlate with volcanic sites or other high silica input sites. Instead it matches the distribution of the silica- secreting organisms. 122

6) The weathering of continental silicate rocks is a continuous process maintaining a permanent flux of oscillating magnitudes of silica into the basins, and thus a pervasive formation of early diagenetic in situ silicified oolites should occur throughout the Phanerozoic where calcareous oolites formed. However, few of the calcareous oolites became silicified (Tomescu et al., 2003), and Phanerozoic onshore-offshore patterns of cherts, as well as the decline of silicified oolites, seems to correlate with distribution patterns of the silica secreting organisms (Kidder and Mumma, 2003). None of the argumenst alone proves the biogenic origin, but integration of all of them is consistent with it.

9.3. Silica burial: a balance between silica dissolution and preservation rates

The geological record of chert deposits is an issue of the preservation rate of biogenic opal that might not correlate with either high biosiliceous productivity areas and high silica-nutrient inputs, or with eutrophic water masses. The net silica output into chert expresses the balance between biogenic silica production, dissolution, dilution by host sediment, and the degree of preservation. A change in the amount of chert burial could be the result of variations in the above factors, and reflect an end product of the conjugated action of the parameters that play a role in the silica cycle and burial as chert. An analysis of the factors that influence the marine biogenic silica production, dissolution and burial is required in order to understand chert abundance patterns through time and their underlying mechanisms. Some key factors include: 1) the productivity of the silica- secreting organisms, 2) the magnitude of the terrigenous sediment input, 3) surface water 123 temperature, 4) water column depth, 5) opal rain rate, 6) the presence of trace metals (Al,

Fe, Ga, Gd, Y), and especially of Al, 7) the presence of organic matter, 8) test aggregation through organism blooms due to high silica-nutrient influxes, 9) the specific surface area of the siliceous skeleton, and burying sediment composition (Nelson et al.,

1995; Maliva and Siever, 1989; Bidle et al., 2002).

The evidence that most modern biogenic silica accumulates in areas with cold surface waters (DeMaster, 1981), suggests that the surface water temperature is a dominant factor for the opal preservation, due to the strong dependence of silica dissolution on temperature (Kamatami, 1982). Based on an actualistic upwelling scenario, biogenic silica accumulations discovered in the stratigraphic record are viewed as tracers of eutrophic water masses induced by the upwelling of cold nutrient-rich deep waters (Ramsey, 1973; Miskell et al., 1985; Hein and Parrish, 1987; Barron and Baldauf,

1995; Murray et al., 1994; Kidder and Erwin, 2001). However, studies showed that the accumulation and burial of modern biogenic silica is governed by different mechanisms in coastal settings and deep-abyssal environments (Nelson et al.1995; DeMaster, 1981;

Tréguer et al., 1995).

• Preservation of silica in the deep-sea: the temperature-dependent preservation model

The deep-sea domain is generally characterized by low terrigenous influx, so that, the sediment input will play a minimal role in the dilution of silica accumulation. Deep- sea sediments are dominated by biogenic components and pelagic clays (Kennett, 1982).

Sediment thickness in the ocean basins depends on the biological productivity, proximity of the basin to areas characterized by high river discharge (hence the amount of sediment 124 influx), rates of dissolution versus rates of preservation of the skeletons, the age of the ocean floor, and the distribution of bottom currents (Kennett, 1982; Duxbury et al.,

2002).

Silica dissolution is strongly temperature dependent, so that the specific dissolution rate (dissolution rate per unit mass of biogenic silica) increases by approximately an order of magnitude with each 15ºC increase in temperature (Kamatani,

1982). Silica dissolution also increases with increasing pressure. Heath (1974) found that a temperature decrease of 1/2ºC per kilometer cancels out the solubility increase due to rising pressure, which accounts for the high silica concentrations are found in deep waters where temperature is low and pressure is high.

In the deep sea, modern biogenic silica accumulations occur in high latitude regions, representing the largest present-day opal accumulation sites (Tréguer et al.,

1995; DeMaster, 1981). The assessment of the abyssal siliceous sediments points out that

“the largest and most important area of modern opal accumulation in the ocean is one where annual rates of opal accumulation in surface waters are not particularly high nor is the annual -based primary productivity” (Nelson et al.,1995; p.366). According to this model, the dominant mechanism for enhancing deep-sea (abyssal) opal preservation is the low dissolution rate of biogenic opal at low surface water temperature (the upper

100 m), irrespective of high or low surface productivity (Nelson and Gordon, 1982 in

Nelson et al., 1995; DeMaster, 1981). By contrast, the oligotrophic areas where the production of biogenic silica in surface waters represents about 75-90% of the global 125 production, is characterized by very little preservation; the dissolution of silica within the water column and at the sediment/water interface is almost total (Nelson et al., 1995).

• Preservation of biogenic silica in coastal margins: the high productivity model

Although the preservation of biogenic opal in abyssal areas is strongly influenced by surface water temperature and latitude, in the shallower waters of continental margins, it is tightly linked to upwelling (Nelson et al., 1995). The upwelling of nutrient-rich deep waters to the favors high organism productivity that increases the opal rain, decreases the test dissolution rate through test aggregation, and leads to enhanced burial rates. The formation of ice caps at the poles triggers low and middle latitudinal deep water upwelling, a mechanical motion that introduces into the nutrient cycling system important quantities of nutrients buried upon deep sedimentation. As Nelson et al. (1995) remark, opal sediment formation in coastal areas is associated with high primary productivity regardless of the latitude.

In view of all of these facts, the analysis of the stratigraphic chert abundance record and fluctuation patterns through time proves to be more complicated than the hypothesized relationships for radiolarian and sponge cherts of high chert abundance = high organic productivity = coastal upwelling of cold nutrient-rich waters, regardless of the silica burial site within the basin, the type of the silica secreting organisms, or the various physico-chemical factors which directly influence silica dissolution and burial rates. Examination of two systems – continental margins vs. deep-water abundance – should reveal distinct differences between these two depth-defined systems that function differently with respect to silica accumulation and preservation (Nelson et al., 1995). 126

However, in the stratigraphic record it is difficult to differentiate between slope and

basinal settings. For this reason, the data of the present study have been divided into very

shallow (peritidal/lagoonal), shallow (shelf), and deep settings (slope/basinal). If biogenic

chert formation in peritidal/lagoonal sites and the shallower parts of the shelf is directly influenced by continental weathering fluctuations via river runoff, that of deeper environments (the deepest area on the shelf, slope and basin) is tightly linked to the intrabasinal ambient (water depth, oceanic circulation, surface water temperature, salinity, seawater chemical composition, volcanic, hydrothermal, and spreading activities, sediment composition) and latitude, in addition to variations of the continental inputs of silica and other nutrients, as well as clastic sediment fluxes. 127

Chapter 10

DISCUSSION: THE ORDOVICIAN SILICA SECRETING BIOTAS - LIVING

STYLE, DISTRIBUTION AND ONSHORE-OFFSHORE EVOLUTION

PATTERNS

The results of Chapter 7 show that at the beginning of the Ordovician the peritidal/lagoonal and shelf waters were inhabited by both siliceous sponges and radiolarians, with the supremacy held by sponges (fig. 10.1). Conversely, the biosiliceous component of slope/basinal sediments was provided mainly by radiolarians, and subordinately by siliceous sponges. During the Middle Ordovician, radiolarians were confined to waters above slope/basin environments, and siliceous sponges retreated from peritidal/lagoonal to shelf and basinal environments, suggesting an Arenigian-

Llanvirnian retreat with a migration time of less than 15 Ma. By the Ashgillian, the chert record suggests that siliceous sponges had retreated completely from peritidal/lagoonal settings, inhabiting mainly shelf environments and subordinately basinal areas which were clearly dominated by radiolarians in terms of biosiliceous accumulation (Fig. 10.1).

10.1. Lower Ordovician

The Lower Ordovician distribution pattern of the dominantly biogenic chert deposits is characterized by shallow water siliceous sponge-dominated cherts contrasting with a predominance of radiolarian chert in slope/basinal settings. This is superimposed 128

Figure 10.1: The Ordovician dominant locus of silica burial and onshore-offshore migration patterns of the silica secreting organisms and the associated facies. The relative size of the radiolarian and sponge symbols reflects the dominance of the corresponding type of organisms in cherts in a particular environment. The relative abundance of the nodular and bedded chert symbols reflects also the dominance of the chert type in a particular environment. Ovals represent nodular chert and horizontal long lines represent bedded cherts. 129

on another distribution pattern whereby shallow water nodular-bedded chert dominates

within carbonate environments, whereas deep water bedded cherts dominate within fine- grained siliciclastic environments (Fig. 10.1). These patterns may represent carryovers from the Late Cambrian and possibly the Middle Cambrian, considering the Middle-

Upper Cambrian bedded chert distribution (Kidder and Mumma, 2003), and the distribution of siliceous sponge and radiolarian faunas (Won and Below, 1999;

Tolmacheva et al., 2001).

Siliceous sponges were able to populate depositional environments of varying depths, from shelf to basinal environments, since the beginning of the Cambrian, when both lithistid demosponges and hexactinellids were already representing a relatively diversified component of the benthic fauna (Finks 1970; Yuan et al., 2002; Li et al.,

1998; Zhang and Pratt, 1994). The fact that the two silica secreting faunas, radiolarians and siliceous sponges, were prevalent in distinct depositional environments emphasizes their ecological separation, probably based on different needs in terms of dissolved silica and other nutrients, water temperature, chemistry, salinity, and sediment substrate.

Considering that siliceous sponges generally need higher dissolved silica levels than the radiolarians (Maldonado et al., 1999), I can hypothesize that during the Lower

Ordovician the shallower siliceous sponge-rich environments, i.e., peritidal/lagoonal and shelf carbonate platforms, were characterized by higher silica levels compared to the surface waters of the deeper settings, a status quite possibly preserved since the

Cambrian. The strontium estimates for the same geological period indeed reveal a relatively high continental radiogenic 87Sr/86Sr input compared to the ensuing Ordovician 130 epochs (Shield et al., 2003) (Fig. 11.1B). Additionally, siliceous sponges are filter- feeding organisms and the very stable, nutrient rich environments offered by the widespread Lower Ordovician carbonate platforms would explain their preference for these shallow habitats.

Studies of Lower Ordovician siliceous sponge assemblages show a differentiation within the siliceous sponge biota itself, namely between demosponges and hexactinellids, in their distributional patterns within a given basin: the lithistid demosponge fauna, quite diversified at that time, prefered the shallow, warm environments offered by the carbonate platform factories, while hexactinellid sponges, much less diversified, occupied deeper environments of slope and basinal sequences within siliciclastic environments

(Mehl and Lehnert, 1997, in Beresi and Esteban, 2004; Beresi and Rigby, 1993, in Beresi and Esteban, 2004; Carrera and Rigby, 1999; Beresi and Esteban, 2004; Tolmacheva et al., 2001). This environmental segregation of the Early Paleozoic siliceous sponge fauna is supported by modern siliceous sponge assemblages. The deep-water Hexactinellids need higher silica concentrations than the shallower water Demosponges (Maldonado et al., 1999). This evidence seems to contradict the hypothesis that during the Lower

Ordovician the shallower marine habitats might have been richer in silica and other nutrients than the deeper areas. However, it suggests that other environmental factors, such as the nature of the sediment substrate and seawater temperature, in conjunction with a high silica level might have been responsible for a dominant species-based distribution and it does not rule out the possibility that there was even more dissolved silica in deeper waters like today. It also suggests that the Early Paleozoic siliceous 131

sponge fauna may have been represented by species different from the modern ones

and/or that the distribution of the Early Paleozoic siliceous sponge fauna was governed

by different factors not present in the modern ecosystems. Overall, during the Lower

Ordovician the Hexactinellid sponges (the deep water population) were a less important

component of sponge faunas compared to the lithistid demosponges (the shallow water

population) that were predominant in the sponge fauna (Finks, 1970; Rigby, 1983;

Carrera and Rigby, 1999). The dominance of Lithistids may explain the abundance of

shallow-warm-water siliceous sponge assemblages within a carbonate environment,

relative to the deep-water siliceous sponge species within a siliciclastic setting, and a

possible biogenic origin via siliceous sponges of the peritidal-shelf nodular cherts and the shallow water bedded cherts (Chapter 9).

More recent findings bring new data relative to the Ordovician siliceous sponge fauna. The shallow-water carbonate sequences found in the Argentine Precordillera show a clear dominance of lithistid demosponges in the siliceous sponge fauna (Beresi and

Esteban, 2004). That study, representing the broadest South American record of

Ordovician siliceous sponge fauna, supports the affinity of lithistid demosponges to warm shallow carbonate environments, for a fauna included in the “eastern North American

Laurentian faunas” (Carrera and Rigby, 1999; Beresi and Esteban, 2004).

Is silica or nutrient limitation of greater importance? This is a question raised by many authors trying to explain the growth and evolution of the siliceous sponges and radiolarians (Maldonado et al., 1999; Maliva et al., 1989). Modern studies on siliceous

sponges still disagree about which parameter exerts the main control, and consequently 132

two antagonistic theories have been proposed: the silicon limitation growth theory

(Reincke and Barthel, 1997; Maldonado et al., 1999; Racki and Cordey, 2000) and the

nutrient limitation growth theory (Frøhlich and Barthel, 1997; Pettijohn, 1975; Siever,

1962) (see Chapter 8).

The silicon limitation growth theory is based on experimental demonstration that the silica uptake rate of siliceous sponges increases asymptotically with increasing silica concentration (Reincke and Barthel, 1997; Maldonado et al., 1999). Insufficient silica concentration weakens the ability of the siliceous sponges to develop complex spicule morphology, and very simplified, small and thin skeletons are secreted instead, but the organisms are still capable to survive. Alternatively, seawater with silica concentrations in excess of 100 µM may inhibit silica uptake (Maldonado et al., 1999). The same relationship was observed in the radiolarians’ behavior: while they are very effective in

silica uptake, occurring abundantly in surface waters that exhibit low silica

concentrations (ca. 1 µM) and yet mineralizing robust siliceous skeletons, silica levels as

high as 100-150 µM have a deleterious effect on their longevity, growth, and weight gain,

acting as a toxin that inhibits their metabolic activity and kills them (Sugiyama and

Anderson, 1997).

Conversely, the nutrition limitation growth theory is based on experimental data showing that silica uptake is an energy-consuming process and therefore in a nutrient- starved environment (as that proposed by Pettijohn, 1975, to explain the inorganic formation of some cherts within a lagoonal setting with restricted circulation and starvation of nutrients other than silica), the biogenic extraction of dissolved silica by 133

siliceous sponges will decrease by 15% (and presumably ceases?) (Frøhlich and Barthel,

1997). Regarding the modern radiolarian fauna, a direct relationship was observed

between their abundance and eutrophic, nutrient-rich, highly saline (40‰) warm (28ºC)

surface waters principally of the equatorial zone or associated to subtropical and tropical

environments (Anderson et al., 1989c). Numerous studies point out that the spatial

distribution of radiolarians might be controlled mainly by water salinity and temperature

(Anderson et al., 1989a,b,c; Anderson et al., 1990; Matsuoko and Anderson, 1992; Chang

et al., 2003), showing a negative correlation with nutrients and primary productivity of

the surface water (Chang et al., 2003). However, regardless of which parameter prevails,

during the Lower Ordovician the strontium isotopic ratios are consistent with the shallow

environments being continuously enriched in silica and other nutrients through riverine runoff, which may account for the high abundance of the more silica-dependent siliceous

sponges (Fig. 10.1).

The dominance of radiolarians in deeper-water environments may be explained

and interpreted in three ways: first, a vestige from the Middle or Upper Cambrian, and

possibly of their deep water origin (Tolmacheva et al., 2001); second, while the deep

water environments are always characterized by an adequate nutrient supply for their

biological needs, other parameters must have prevailed in stabilizing the dominant

basinal distribution of radiolarians; and third, they thrive principally in upwelling settings

where silica and other nutrients are readily available.

The paleolatitudinal distribution of Lower Ordovician cherts reveals a low to mid

latitudinal increase in silica burial (Fig. 7.4 and Table 7.5) with the radiolarian dominated 134

deep-water and peritidal/lagoonal deposits confined principally to low paleolatitudes (0° -

30°), and the mid-latitudinal cherts (30° – 60°) occurring mainly in shelf deposits (North

China and Korea). However, relative to the Lower Ordovician paleogeography of Korea and North China, more recent paleogeographic studies (Xiaofeng and Xiaohong, 1995) place the benthic faunas at low latitudes and not at mid latitudes as shown in Fig. 4.

The distribution analysis of the chert formations bearing biogenic siliceous remains shows their low latitudinal expansion (0° - 30°) was principally in the southern hemisphere, in Kazakhstan, South China, and the southern margin of Laurentia. The reports from Bolivia, Poland, Korea, and North China do not include any mid-latitudinal cherts that contain sponge or radiolarian fragments, but more recent findings of studies from northern Argentina which was situated at middle paleolatitudes (Fig. 7.4) demonstrate the presence of hexactinellids in siliciclastic sedimentary sequences during the Lower Ordovician (Beresi and Esteban, 2004). However, it is worth mentioning that the paleogeography of the Argentine Precordillera (western Argentina) is a debated problem: some workers consider that it was a part of Gondwana during the Cambrian, and that it collided with Laurentia during the Ordovician (Dalla Salda et al., 1992), while others argue that it was a microcontinent that rafted from the Ouachita embayment of

Laurentia during the Cambrian, and collided with Gondwana during the Middle

Ordovician (Astini, 1995; Thomas and Astini, 1996). The latter view seems to be

supported by more pertinent evidence and is more widely accepted.

Assessing this global distribution pattern and interconnecting multiple factors

related to it, a paleolatitudinal based partition is evident: 1) deep-water, radiolarian 135 dominated cherts are present in low latitudes; 2) lithistid demosponge dominated chert facies dominate peritidal/lagoonal and carbonate shelf; and 3) the mid latitudes are characterized by deep shelf and basinal environments dominated by Hexactinellids.

The confinement of radiolarians to low latitudes offers insights into some ecological aspects of the Early Paleozoic radiolarian fauna. Their apparent preference for warmer waters, was a consequence of since their emergence in the greenhouse Middle

Cambrian climate (De Wever et al., 2001). The mid and high latitude Lower Ordovician waters were probably too cool for their survival, suggesting that during that time there was a latitudinal thermal gradient high enough to constrain radiolarian latitudinal distribution. The distribution of Early Paleozoic radiolarians is tightly linked to water temperature, a hypothesis suggested by Sugiyama and Anderson (1997) for modern radiolarian communities, and in agreement with the general line of thought that associates plankton distribution with abiotic factors such as temperature and latitudinal zonation

(Lipps, 1970; Tappan and Loeblich, 1973).

Siliceous sponge faunas are able to inhabit environments with a wider latitudinal distribution probably because of: 1) either the specialization of the two different sponge groups, Lithistid Demosponges and Hexactinellids, to life in different environments, shallow warm carbonate settings for the former, and respectively cooler, deeper, siliciclastic habitats for the latter, perhaps an evolutionary ability gained due to their participation in and survival from the Latest Precambrian Icehouse; or 2) they evolved as primitive generalists that could adapt readily to variable environments as long as they have clear waters and sufficient nutrients and dissolved silica. 136

10.2 Middle Ordovician

The distribution of Middle Ordovician cherts within siliceous skeletal remains

reveals that radiolarians are exclusively confined to slope/basinal environments whereas

siliceous sponges seem to characterize both shelf and slope/basinal settings (Fig. 10.1).

The Middle Ordovician distribution pattern of silica secreting biotas reflects a change in of siliceous sponge faunas from peritidal/lagoonal to shelf-basinal environments, and of radiolarians to basinal settings, suggesting an Arenigian- Llanvirnian onshore-offshore migration of the silica secreting organisms in less than 15 My. This shift of the cherts containing preserved skeletal remains is accompanied by a general onshore-offshore drift in the locus of marine silica deposition, so that during the Middle Ordovician the slope/basinal environments represent the dominant domain of silica burial. The faunal retreat is also coeval with a turnover in the chert style from shallow-water carbonate environments dominated by nodular-bedded cherts characteristic for the Lower

Ordovician, toward a slope/basin bedded chert-dominated silica system within siliciclastic environments in the Middle Ordovician (Fig. 10.1).

10.2.1. Migration mechanisms

Speculations can be made about the style of the organism migration. If

radiolarians moved offshore because of various causes, if the linkage between

radiolarians, a phytoplankton food source, and siliceous sponges, a sessile benthic component of the food web, was as strong as in modern marine communities, and if silica

was a limiting factor for the Ordovician siliceous sponge fauna, siliceous sponges may 137

have followed the offshore migration of radiolarians (Kidder, 2000). Today, radiolarians and sponges are both consumers but at different trophic levels (De Wever et al., 2001).

The higher trophic levels characteristic for sponges depend on the lower levels including radiolarians. Within the food web as we know it today, benthic faunas, especially sessile organisms such as sponges, depend on the more flexible planktonic organisms. Siliceous sponges may have followed radiolarians because the marine environments proximal to those inhabited by radiolarians were richer in dissolved silica owing to the fact that radiolarians dissolve at faster rates than siliceous sponges (Scholl et al1983).

Potential mechanisms for the shallow to deeper water migration of the silica secreting biotas include: i) onshore-offshore evolutionary patterns, and/or ii) changes in availability of dissolved silica and other nutrients used by siliceous sponges and radiolarians (Tomescu and Kidder, 2002).

i) Onshore-offshore evolutionary patterns of silica secreting organisms

The shift of cherty facies from nearshore settings toward deeper environments might have been driven by onshore-offshore patterns of evolution (Kidder, 2000), an

Ordovician mechanism suggested for benthic invertebrate calcareous macrofaunas

(Sepkoski and Sheehan, 1983; Jablonski et al., 1983). The Ordovician contains evidence of an onshore-offshore migration of the benthic communities of the three Phanerozoic faunas: the declining Cambrian fauna became restricted more or less to deep water environments moving offshore perhaps by the Paleozoic fauna diversifying across the shelf, and the Modern fauna developing in nearshore environments. The displacement of the Cambrian fauna by the Paleozoic fauna occurred over the Lower and Middle 138

Ordovician, and the rise of the Modern fauna in nearshore settings occurred in the mid to late Ordovician (Jablonski et al., 1983; Sepkoski and Miller, 1985). However, these

analyses of Early Paleozoic marine faunas do not include any planktonic group of

organisms, and it is unclear whether they apply to the siliceous sponges. Camparable

timing of migration would be consistent with siliceous biotas moving in concert with

calcareous ones.

ii) Changes in availability of dissolved silica and other nutrients

The onshore-offshore migration of the silica secreting biotas can be explained in

terms of a shift in the dominant source of silica and other nutrients (P, N, Fe), from an

allochtonous-dominated (exogenous input through riverine runoff) to an autochtonous-

dominated (endogenous inputs through volcano-hydrothermal and seafloor weathering

activities) system.

In modern oceans, among the four silica sources that replenish the marine silicic

acid, river runoff accounts for 82%, volcano-hydrothermal together with seafloor

weathering inputs only 10% of the total silica, and the remaining of 8% is attributed to

eolian supply (Tréguer et al., 1995). The nutrient load of the rivers is strongly influenced

by the silicate weathering rates and the volume of rocks exposed to weathering.

Consequently, a change in weathering will affect the river flux, that in turn induces

changes in the marine realm. The volume of silicate rocks exposed to weathering, and the weathering rates are directly related to changes in uplift of mountain belts, climate, sea level fluctuations, the interplay between the latitudinal disposition of the landmasses and 139

the emerged land area, and indirectly related to tectonic rifting and oceanic spreading

rates.

The Ordovician was a geological period characterized by energetic mid-oceanic

ridge activity (Mackenzie and Piggot, 1981), intensive volcanism (Ronov, 1982; Huff and

Kolata, 1990; Huff et al., 1992; Huff et al., 1996) common in submarine islands, coastal

and cordilleran sites (Stillman, 1984), numerous volcanogenic massive sulfide deposits

compared to the Cambrian and the Silurian (Eastoe and Gustin, 1996), a low area of

continental deformation and orogenic activity during the Lower and Middle Ordovician

(Richter et al., 1992; Miller and Mao, 1995), and intensive reorganization of the major

tectonic plates (Ford and Golonka, 2003; Worsley et al., 1994).

The 87Sr/86Sr ratio in marine sediments is tightly linked to the balance between continental weathering (radiogenic flux) and oceanic spreading rates (non-radiogenic flux). The minimal change in the ration through the Lower and Middle Ordovician is consistent with a relatively unchanging chemical dissolved runoff from the landmasses.

10.2.3. Considerations regarding the Middle Ordovician sea level fluctuations

The Middle Ordovician is a geological time characterized by a generally much lower sea level compared to the Lower Ordovician and the ensuing Caradocian (Ross and

Ross, 1992). The sedimentary sequences of most of the Whiterockian series (the lower part of the second-order Tippecanoe sequence of the North American stratigraphy) reveal a sharp Middle Ordovician sea level drop that exposed almost the entire craton of

Laurentia (Finney, 1997; Fortey, 1984). The evidence is based on both deep 140

(slope/basinal) (Finney, 1997) and cratonic shelves sedimentary successions (Ross and

Ross, 1992; 1995). The Whiterockian – Ibexian boundary in North America coincides

approximately with the Lower – Middle Ordovician boundary (late Arenigian, Fig. 6.2).

According to Finney (1997), this regression boundary is a second-order eustatic sea level

drop that represents a conspicuous horizon recognizable across the North American continent, and which lasted 3 to 5 million years. The Sauk-Tippecanoe sequence boundary, known as a global paleokarst unconformity (Sloss, 1963), has been also recognized in South Korea, North China, the Canadian Appalachians, and the Southern

Appalachians (Ryu, 2002), in southern Wales, on the Australian platform, as well as in

Sweden (Fortey, 1984), and northern Europe (Webby, 1984a,b).

A sea level drop results in a reduced habitable space on the shelf and thus shallow water communities would be forced to track the drop of their habitat (Fortey, 1984).

However, shallow waters would still exist even though these environments would suffer the greatest loss in space, and coastal settings might not change much in areal extent. The degree to which sea level changes and glaciations affect is controversial (e.g.,

Wilde and Berry, 1984; Fortey, 1984).

Unfortunately the potential mechanisms responsible for the global early Middle

Ordovician sea level drop are not well documented. According to Ryu (2002) regional

and tectonic events generated local and regional variations of the Sauk-Tippecanoe

sequence boundary. Considering that the Sauk-Tippecanoe sea level drop is seen as a

major eustatic event, driving mechanisms might be represented by: 141

1) fluctuations in the rate of sea-floor spreading that would directly disturb

significantly the volume of ocean water (Donovan and Jones, 1981; Hays and Pitman,

1973). This mechanism that would be in agreement with a decrease in the rate of seafloor

spreading suggested by François et al. (1993, after Gaffin, 1987), and implicitly a

decreasing area of continental deformation (Richter et al., 1992; Qing et al., 1998). A

decrease in the volume of oceanic ridges would affect directly the basin volume process

depending on oceanic spreading rates (see paragraph 1) – but seen as a long-term

mechanism, lasting 10-30 m.y. (Donovan and Jones, 1981).

2) changes in mean oceanic temperature, through thermal seawater contraction;

these would have an effect of smaller magnitude susceptible to be masked by the effects

of other processes (Donovan and Jones, 1981).

3) increase in the volume of ice sheets (Fortey, 1984; Lindström, 1984) based on

the geographic position of Gondwana over the South pole. However, no glacial deposits

support this hypothesis.

10.3. Upper Ordovician

The Upper Ordovician distribution of cherts illustrates a depositional environment

distribution pattern similar to that of the Middle Ordovician and consistent with the proposed onshore-offshore patterns inferred previously. Radiolarian bedded cherts were confined to slope/basinal settings, and shelf deposits are dominated by nodular and bedded cherts with siliceous sponge spicules preserved. The three Caradocian silicified oolites on the southern margin of Laurentia suggest that siliceous sponges might have 142

still been present sporadically in some peritidal/lagoonal areas characterized by peculiar

local conditions. During the Ashgillian, the peritidal/lagoonal cherts disappear

completely. If the chert depositional migration occurred within the same range of low and

middle paleolatitudes in the Lower and Middle Ordovician, during the Upper Ordovician it expanded to the high paleolatitudinal habitats, (the nodular cherts from Italy and

France; Loi and Dabard, 2002). 143

Chapter 11

INTERPRETING THE CHERT ABUNDANCE PATTERNS: WHY CHERT

PEAKS OR LOWS?

11.1. Lower Ordovician: a greenhouse interval

The present study demonstrates that the Lower Ordovician was characterized by high chert abundance principally in shallow, peritidal/lagoonal, and shelf environments and secondarily in slope/basinal settings (fig. 7.1A, 7.3, 10.1, and 11.1A). The Lower

Ordovician was a period characterized by: 1) high sea level, including the climax of the

Precambrian – Lower Ordovician Sauk transgressive sequence that occurred during the late Ibexian (early Arenigian; fig. 6.2), when the sea level reached one of the highest levels in the entire Phanerozoic (Finney, 1997; Ford and Golonka, 2003); 2) maximum dispersion (disassembly) of continents by the northward drifting of Laurentia and Baltica concomitant with widening of the Iapetus Ocean and Tornquist Sea (Worsley et al.,

1994; Ford and Golonka, 2003); 3) intensive volcanic activity since the late Cambrian

(Stillman, 1984); 4) widespread carbonate platforms (Overstreet et al., 2003); and 5) organism radiation (fig. 11.1A). The strontium curve (fig. 11.1B), during the early

Lower Ordovician, that is the Tremadocian, has higher radiogenic values than during the ensuing Ordovician epochs, suggesting higher continental weathering rates of the silicate rocks exposed by the Pan-African orogeny (Ebneth et al., 2001), but still low 144

Figure 11.1A: The Ordovician biotic system 145

Figure 11.1A – The Ordovician biotic system: continued. a. Chert abundance patterns evidenced by the present study. The data is normalized to time. b. Chert abundance by depositional environments, as evidenced by the present study. Gray bullets represent peritidal/lagoonal cherts, black bullets represent shelf chert deposits, and stars represent slope/basinal chert formations. The data is normalized to time. c. Radiolarian diversity; after Vishnevskaya, 1997. d. Black shale abundance patterns in the British Isles; after Leggett et al., 1981. e. Euxinification episodes based on black shale abundance; after Leggett et al., 1981. f. Graptoloid diversity; after Leggett et al., 1981. g. diversity; after Leggett et al., 1981. 146

Figure 11.1B: The Ordovician abiotic system 147

Figure 11.1B – The Ordovician abiotic system: continued. a. Strontium and oxygen isotope curves based on calcitic ; after Shields et al. (2003). b. Extent of continental deformation areas; after Reichter et al. (1992) and Qing et al. (1998). c. Sea level fluctuations; after Ross and Ross (1992; 1995); Qing et al. (1998); Finney, (1997). d. Tropical seawater temperature based on oxygen isotope values; after Shields et al. (2003) and Tobin et al. (1996). e. Tectonic data: orogenies and continental migration patterns; duration of Taconic orogeny and continental migration patterns after Ford and Golonka (2003), Qing et al. (1998). A Ringworld is dominated by tropical continents, while a Capworld is characterized by polar continents (Worsley and Kidder, 1991; Worsley et al., 1994). Question marks for uncertain timing. f. Average latitude of the world landmass for the “2-10” model (Worsley and Kidder, 1991; Worsley et al., 1994). g. Volcanogenic massive sulfide deposits (Eastoe and Gustin, 1996). h. Oolitic ironstone accumulations (Van Houten and Arthur, 1989). 148 enough to allow the development of widespread carbonate platforms, that were overwhelming the shelf system, up to relatively high latitudes (50º in Asia; Frakes et al.,

1992). Shelf areas were very extensively developed in North America, Australia, Asia, and South America, and were centered close to the equator and tropics. The widespread marine carbonates deposited in shallow epeiric seas (Overstreet et al., 2003), the oxygen isotopic data (Tobin et al., 1996; Shields et al., 2003), and the relatively abundant oolitic ironstone deposits (Van Houten and Arthur, 1989) provide support for a warm and humid greenhouse mode.

11.1.1. Peritidal/lagoonal and shelf environments

The peritidal-lagoonal and shelf environments during the Lower Ordovician were inhabited mainly by siliceous sponges and subordinately by radiolarians (fig. 10.1), resulting in abundant nodular and less frequent bedded cherts (fig. 7.3 & 7.7).

Geographically, approximately 75% of the worldwide Lower Ordovician shallow water cherts have been identified on the south-southwestern margin of Laurentia, such as on the Appalachian and Cordilleran passive margins, the south Oklahoma aulacogen, and the Ozark Uplift (Overstreet et al., 2003) (Table 7.6), as well as in the Kazakhstan block,

China, Korea, Australia, and Argentina (fig. 7.4).

• Why abundant chert burial?

The modern silica cycle does not exhibit any shallow biogenic opal accumulation and burial, and therefore we can only speculate about the driving mechanisms that generated biogenically-induced chert formation in the past. First of all, both peritidal/lagoonal and shelf depositional environments were inhabited by silica secreting 149 organisms, principally siliceous sponges and secondarily radiolarians, which provide the necessary silica for the chert formation. Also, these shallow water settings were characterized by a generally low terrigenous input that could not mask the biogenic silica accumulation, or inhibit the development of large carbonate platform factories, and the expansion of filter feeding siliceous sponges and other benthic organisms. Jones and

Desrochers (1992) emphasize that shallow platform carbonates can form in areas isolated from extensive siliciclastic flux. The strontium curve (fig. 11.1B) is consistent with the fact that Lower Ordovician riverine runoff must have been still high enough to account for the needs in terms of dissolved silica and other nutrients required by siliceous sponges and radiolarians. Moreover, the host rock represented by carbonates probably played an important role in the chert formation and burial, based on the affinity between carbonates and chertification (Kastner et al., 1977; Maliva and Siever, 1989). Another factor in chert formation could be the presence of abundant organic matter whose decomposition through oxidation may induce early diagenetic nodule formation (Siever,

1962; Maliva and Siever, 1989). According to the force of crystallization-controlled replacement model of Maliva and Siever (1989), the ideal host sediment has three properties: organic matter content, porosity and permeability, and biogenic-opal concentration.

Taking into account that seawater was warm (as shown by oxygen isotopic ratios; fig. 11.1B), the opal dissolution rate is should have been high and thus driving silica recycling and preventing opal accumulation. However, the loss of opal through dissolution would have been counterbalanced and surpassed by: 1) the dominance of 150

siliceous sponges that dissolve at lower rates than radiolarians (Scholle et al., 1983); 2)

the shallow depth of the water column that would reduce the time available for sinking

and dissolution of opal radiolarian skeletons; 3) low terrigenous input that could mask

silica concentrations; 4) the carbonate composition of the host rock; 5) the presence of

abundant organic matter, and 6) a possibly rapid burial of the benthic silicisponges

(before their silica can get into ocean water) through a relatively high vertical

accumulation of the carbonate production; instead, the dissolution of the opaline skeleton

makes chert nodules.

In conclusion, while we can only speculate about a good organism productivity in

these shallow settings, due to stable and favorable environmental conditions, the chert

abundance could be accounted for in a low sediment dilution - enhanced chert formation

model (fig. 11.2).

11.1.2. Slope/Basinal environments

The slope/basinal settings were inhabited principally by the planktonic radiolarians, and subordinately by siliceous sponges. It was demonstrated that at least in certain basinal environments the bottom waters were well aerated (Wilde, 1989) and could sustain the activity of some benthic siliceous sponges (Blome et al., 1995a;

Tolmacheva et al., 2001). Also, Leggett et al. (1981) discussed the presence of burrows and primary iron in cherts of the British Isles (Iapetus Ocean) associated with pillow lava, that indicate well oxygenated bottom waters, as well as high silica and other nutrients inputs. The well oxygenated deep water theory is also supported by the presence of deep burrowing trace fossils (up to 79 cm deep), discovered in an early Ordovicia 151

Figure 11.2: Chert abundance Model. LOWER ORDOVICIAN [490 – 470 Ma] 152

deep-water flysch sedimentary sequence of Nova Scotia (the Meguma Group; Pickerill

and Williams, 1989). That study represents the first evidence of infaunal behavior in deep water environments, in contrast to the previous much shallower estimates that were based on the analysis of shelf settings (Pickerill and Williams, 1989). Also, according to Crimes

(1974; 1992) a clear increase in the deep-water ichnofacies can be seen during the Ordovician.

On the other hand, a warm world with warm surface waters would lead to higher metabolic activity, higher silica dissolution and thus higher silica recycling in the uppermost water layers. This recycling would lead to a decreased potential for biogenic opal preservation. The significant silica accumulation and burial in the deep-sea may account for high preservation rates at the bottom part of the basin. If under these circumstances additional processes generating silica and other nutrients inputs (e.g., volcanic, hydrothermal, and oceanic spreading activities) were active, these could create an environment with a higher potential for blooms of silica secreting organisms, higher opal rain rates, and a higher potential to increase the silica net output through the formation of biogenic chert.

A relatively high radiolarian productivity within a warm water world is consistent

with: 1) the ecology of some modern radiolarians shown to be tightly linked to warm,

highly saline surface waters (Anderson et al., 1989c); 2) the origin of radiolarians as part

of the Cambrian explosion that occurred in a warm climate; 3) the theory that submarine

volcanism and hydrothermal activity, coupled with a warm climate, and a sea level rise,

provide the right conditions for increased productivity (Vermeij, 1995); and 4) major 153 tropical continental margins upwelling sites along the equator in western America and

Arctic Islands in Laurentia, North China, New Zealand, and Newfoundland (Wilde,

1989).

In concordance with Vermeij’s theory (1995), a high temperature constitutes the main stimulus for high productivity through enhanced metabolico-chemical reactions that are greatly accelerated with increased temperature. The model emphasizes the fact that increased nutrient supply only is not enough to generate significant high productivity, and that organisms revolutions are mainly generated by increased energy

(such as temperature) and subordinately by nutrient supply. The survival of radiolarians during this interval, along with observations showing that for modern taxa high temperatures above 33º C have deleterious effects on radiolarian longevity (Anderson et al., 1987c), indicate the presence of favorable marine environments. The presence of abundant radiolarians appears to rule out the possibility of overly high surface seawater temperatures at low latitudes, which are suggested by the very negative oxygen values obtained for this epoch (fig. 11.1B; Shields et al., 2003; p. 2016).

Good preservation rates within a well-aerated deep bottom environment may be the result of the interplay of multiple factors such as: 1) a fairly high rate of opal rain induced by a good productivity in the surface waters, 2) the presence of organic matter that intensifies silica preservation (Spencer, 1983; Siever, 1957); 3) a water column shorter than that of the modern abyssal sites directly dependent on the basin morphology, and 4) possibly higher concentrations of cations, such as Al, Fe. These cations that were shown to combine with diatom silica to decrease dissolution rates (Lewin, 1961). Nelson 154 et al., (1995) assert the dissolution rate and the equilibrium solubility of diatom silica decreases as the Al/Si ratio increases.

The clustering of Lower Ordovician basinal bedded cherts on the southeastern margin of Laurentia, might be because of least two of the factors that would enhance silica preservations: 1) high cation concentrations of intrabasinal origin, through submarine volcanism driven by the widening of the Iapetus Ocean (Ford and Golonka,

2003) and 2) favorable organic matter concentrations, such as the presence of zooplanktonic graptoloids whose diversity reaches the highest peak for the Ordovician, at least in the Iapetus Ocean (fig. 11.1A; Leggett et al., 1981). Even though black shale studies do not indicate any significant organic-rich sediments formed during this interval, a finding otherwise consistent with well aerated bottom demonstrated for the

Iapetus Ocean (fig. 11.1A; Leggett et al., 1981), the levels of organic matter were seemingly high enough to enhance silica burial in concert with other factors. Though it is difficult to prove or test, a preservation model is consistent with abundant chert in a green house model.

However, there is no study showing a relationship between radiolarians and Al intake similar to the one documented for diatoms. Also, the role of organic matter in silica preservation is arguable. It may be enhancing by its adsorption on the silica surface or, on the contrary it may be deleterious by the formation of organic-silica complexes, so that no conclusive relationship has been supported to date (Siever, 1962). Bidle and

Azam (1999) and Bidle et al., (2002) showed that bacterial activity is important in silica 155 dissolution by degrading the organic film that protects the opal diatomaceous frustules from dissolution.

In conclusion, the dominance of bedded cherts in basinal settings (fig. 7.7A) might reflect relatively high biosiliceous productivity, associated with good taphonomic conditions (high preservation rates) induced by the conjunction of various opal enhancing mechanisms, under an oxic regime within the bottom areas (fig. 11.2).

11.2. Middle Ordovician: a greenhouse interval (fig. 11.3)

The general chert abundance is strongly diminished during the Middle

Ordovician, exclusively due to a 75% decrease in silica burial in peritidal/lagoonal and shelf environments (Table 7.3 & fig. 7.3). The data show also that the formation of nodular cherts has been intensely affected in comparison with that of the bedded cherts, during this interval (fig. 7.2A). All these facts, along with the genetic relationship that exists between siliceous sponges and peritidal/lagoonal-shelf chert formation (Chapter 7 and 9) suggest that the benthic siliceous sponge biota was more involved in this general chert abundance low. The slope/basinal bedded cherts, during the same interval on the contrary, record a slight augmentation suggesting that radiolarian faunas and implicitly radiolarian-induced chert formation were not responsible for the Middle Ordovician global silica shortage. Geographically, the most affected area by the silica burial low was

Laurentia, which suffered a 50% reduction in chert formation (Table 7.6).

The Middle Ordovician shallow water chert low may reflect either an apparent, or a real silica burial slowdown. The apparent curtailment may be due to some common 156

Figure 11.3: Chert abundance Model. MIDDLE ORDOVICIAN [470 – 458 Ma] 157

errors and biases inherent to any compilation study, such as sampling issues, afe assignment problems, and/or to the fact that the worldwide sedimentary sequences are unequally studied so that the absence of chert deposits in certain geographical areas might reflect just a lack of knowledge and not a physical absence.

On the other hand, an actual slowdown of silica burial may be linked to changes

in the biotic and abiotic Middle Ordovician systems. The Lower-Middle Ordovician

onshore-offshore organism migration, especially that of siliceous sponges, that have

been demonstrated herein to be responsible for the formation of shallow chert deposits

and which proved to be most affected by this chert waning. Physico-chemical changes of

the depositional environments (e.g., a sea level transgression) that would have influence

the balance between opal dissolution and opal preservation and the balance between

shallow (peritidal/lagoonal and shelf) and deep water (slope/basinal) environments.

The Middle Ordovician represents a time of (fig. 11.1): 1) onshore-offshore

migration of both benthic and planktonic organisms that try occupy new ecological

niches (Sepkoski and Sheehan, 1983; Jablonski et al., 1983); 2) major plate tectonic

reorganization as the Iapetus Ocean and Tornquist Sea begin to narrow down, and the

Rheic Ocean is forming (Ford and Golonka, 2003); 3) low relative sea level (Ross and

Ross, 1995; Finney, 1997); 4) warm surface water temperatures (27ºC - 32ºC; Shields et

al., 2003); 5) decreased emergent land area coeval with lower average latitude of the

world landmasses (Worsley et al., 1994); 6) higher radiolarian diversity (Vishnevskaya,

1997); and 7) more abundant back shale deposits (Leggett et al., 1981). However, the

present study reveals a decrease in the mid latitudinal cherts concomitant with an 158 increase of the low latitudinal deposits (Table 7.5). That suggests either a less warm climate compared with that of the Lower Ordovician or less available shelf area in which chert could form.

11.2.1. Peritidal/lagoonal and shelf environments

The relationship between the formation of shallow water cherts and siliceous sponges (Chapter 9), and the fact that Middle Ordovician proves to be a geological interval of biological adaptation to new, deeper-water ecological niches by a onshore- offshore organism migration (see Chapter 10), suggest that the Middle Ordovician chert low was driven by a decrease in the shallow water silica burial due to unstable conditions in the evolutionary scenario of siliceous sponges and subsequently to a lesser extent of radiolarians.

Another mechanism that could impact the Middle Ordovician chert preservation is represented by a reduction in the shelf area in which cherts could form. In the early

Middle Ordovician, the main Lower Ordovician chert bearing continent, Laurentia, was entirely emergent due to a short-lived major regression (Finney, 1997) recognizable worldwide (Ryu, 2002; Fortey, 1984; Webby, 1984a,b; Sloss 1963) (Chapter 10). Thus, it can be argued that the slowdown in chert deposition may be strictly a consequence of the physical displacement of shelf environments by the retreat of the sea level, and not linked at all to the organism migration. Finney (1997) estimated that the regression had a short duration (3-5 Ma), and that the rest of the Middle Ordovician and up to the early

Caradocian witnessed a gradual transgression and implicitly a re-expansion of the shelf area. However, at this point in time, the siliceous sponges inhabited deeper sites of the 159

carbonate shelf environment and radiolarians were populating exclusively the

slope/basinal waters, so that siliceous sponges could not compensate for the drop in

overall shallow water chert production.

11.2.2. Slope/basinal environments

The planktonic organisms (e.g., radiolarians and graptoloids) that accumulated in

deep waters would have greatly benefited from a higher intrabasinal input of silica and

other nutrients (P, N, Fe). Such an increased input is consistent with the strontium curve

(a change that has occurred before the Lower-Middle Ordovician boundary, sometime in

the Ashgillian), and is due to an energetic midocean activity and a higher infaunal activity resulting from the invasion of deeper nutrient-rich areas as a consequence of the

onshore-offshore Ashgillian-Llanvirnian organism migration. Innovations in burrowing

organisms such as the population of new deeper-water ecological niches would constitute

a stimulus for a higher productivity of the planktonic organisms, by making available

new amounts of nutrients entrapped into sediments during precedent periods. The release

of phosphorous entrapped into the deep-sea sediments triggered by a migration of the

benthic organisms was proposed as one mechanism responsible for the Cambrian

explosion (Bengston, 1994). The increased deposition of black shales (fig. 11.1A) during

this period, coeval with warm surface waters, and a slight increase in the deep basinal

silica burial (fig. 7.3), provide support for good organism productivity (Vermeij, 1995)

concurrent with enhanced silica preservation mechanisms (i.e., a higher influx of organic

matter due to a higher productivity). The intense deep water activity of the benthic faunas 160 indicates the presence of oxic bottom deep waters at least in some regions (Leggett et al.,

1981). However, the higher abundance of organic-rich sediments intercepted through more abundant black shales compared to that of the precedent period (fig. 11.1A), suggests a higher organic productivity and/or somewhat less oxygenated bottom environment than that of the Lower Ordovician. Consequently, the high basinal chert abundance during this interval seems to conform to a high organic productivity and enhanced preservation rates model (fig. 11.3).

11.3. Upper Ordovician: the Caradocian - a warm - cool transitional, ice free

mode (?)

During the Caradocian, opal burial reached the highest Ordovician peak in both shelf and slope/basinal environments. A balance between shelf and deep-water cherts that was tipped heavily toward deep-water cherts during the Middle Ordovician is restored. While radiolarians are confined to deep basinal cherts, siliceous sponges populate principally carbonate shelf areas and secondarily deep water niches (fig. 10.1).

The three Caradocian silicified oolite occurrences on the southern margin of Laurentia suggest that siliceous sponges may have still been present sporadically in some peritidal/lagoonal settings that were characterized by peculiar local conditions.

Geographically, approximately 74% of the worldwide Caradocian cherts have been found around Laurentia, mainly on the southwestern, southern, and southeastern margins. The remaining occurs in Australia and China (23%), and on the Kazakhstan block (3%). Overall, the high chert abundance suggests a biotic-abiotic earth system 161

highly favorable for chert burial. The changes in the rate of biogenic silica burial may be

due to fluctuations in biogenic silica production, dissolution, and/or preservation rates

that could be linked to physico-chemical modifications of the Earth system.

11.3.1. The abiotic system

The Cardocian epoch witnessed major changes in abiotic components of the Earth

system. Major physical changes that occurred over the Middle – Upper Ordovician

transition and during the Caradocian includ (fig. 11.1B): 1) a volcanic climax (Stillman,

1984); 2) West Gondwana crossed the South Pole (Ford and Golonka, 2003); 3) the

Taconic orogeny (Ford and Golonka, 2003; Qing et al., 1998); 4) a positive shifts in δ18O

(Shields et al., 2003; Tobin et al., 1996) and δ13C (Patzkowsky et al., 1997) values; 5) a

highly non-radiogenic strontium ratio indicating strong mid-oceanic ridge activity (Qing

et al., 1998; Shields et al., 2003) and large ridge volume (Mackenzie and Piggott, 1981) ;

6) a minimum in continental deformation (Richter et al., 1992); 7) a peak in global sea

level (Ross and Ross, 1992; Qing et al., 1998); 8) abundant volcanogenic massive sulfide

deposits in the Iapetus Ocean (Eastoe and Gustin, 1996); 9) abundant oolitic ironstone

accumulations (Van Houten and Arthur, 1989); and 10) reduced emergent area coeval

with lower average latitude of the world landmass (Worsley et al., 1994) (fig. 11.1B). All

these activities induced changes in the global climate and oceanic seawater that would

affect the biotic system and the biogenically-controlled silica cycle.

One of the most important changes affecting the Upper Ordovician world is the

initiation of climatic cooling and the development of ice sheets over both the South and the North Pole (Poussart et al., 1999), but the timing of these processes, Caradocian 162

versus Ashgillian is still open to controversy (Frakes et al., 1992; Brenchley et al, 1994;

Pope and Steffen, 2003). Although the initiation of permanent ice cap sheets in the

Caradocian is viewed skeptically, it is accepted that the cooling mode that heralded the

late Upper Ordovician glaciation originated in the Caradocian (Patzkowsky and Holland,

1993; Worsley et al., 1994; Tobin et al, 1996; Patzkowsky et al., 1997; Pope and Steffen,

2003; Shields et al., 2003). Recent modeling studies relative to the Late Ordovician

glaciation show the necessary conditions for the formation of ice sheets (Herrmann et al.,

2004). Based on the relationship between atmospheric pCO2 values, sea level, poleward

ocean heat transport, and paleogeography, the simulations show that the high sea level

and atmospheric pCO2 ( 8 x PAL) during the Caradocian could have induced the

formation of small ice sheets if the poleward ocean heat transport was reduced to half of

the modern value. However, Caradocian simulations evidence ice volumes that are strongly inferior to those obtained for Asghillian simulations.

A cooling trend was initiated sometime in the Caradocian based on paleontologic, tectonic and isotopic evidence (Frakes et al., 1992; Patzkowsky and Holland, 1993;

Worsley et al., 1994; Tobin et al, 1996; Patzkowsky et al., 1997; Pancost et al., 1999;

Ainsaar et al., 1999; Pope and Steffen, 2003; Shields et al., 2003). However, the time of

origination of the Caradocian cooling mode is disputed and ranges between the beginning

of the Caradocian, based on oxygen isotopic data (e.g., Shields et al., 2003), and middle-

late Caradocian, based on high planktonic poductivity and carbon isotopic data

(Patzkowsky et al., 1997) and an offset feedback induced by the Taconic orogeny that started earlier (Kump et al., 1999). 163

Oxygen isotopic data and seawater temperature

Numerous studies have looked at the Ordovician seawater oxygen isotopic

composition in brachiopod shells and marine cements such as fibrous calcite. The values

obtained are controversial but all of them show generally a steady rise of the calcite δ18O levels throughout the Ordovician (Tobin et al., 1996; Shields et al., 2003), interpreted as a general cooling trend. Upper Ordovician seawater δ18O values obtained from

geographically widespread stratigraphic sequences (Shields et al., 2003) show an oxygen

depletion trend, by approximately 3‰ PDB that occurred over a period of less than 10

My around the Middle-Upper Ordovician boundary, (fig. 11.1B; Shields et al., 2003).

Two driving mechanisms have been hypothesized with the aim to explain this positive shift in δ 18O (fig. 11.4).

1) The first mechanism is a decrease in minimal tropical seawater temperature

from a 27°-32°C to 16°-25° C, within an ice-cap free world, for a δ18O SMOW (standard

mean oceanic water) of -3‰ (Frakes et al., 1992; Tobin et al., 1996; Shields et al., 2003)

(fig. 11.1B and 11.4). The estimate for the temperature drop is in agreement with that of

Worsley et al. (1994) for equatorial temperatures as low as 25ºC, a cooling considered as

having occurred prior to the late Ordovician glaciation and within a world that lacked ice

sheets.

2) The second mechanism is given by a decrease of the δ18O SMOW by 2-3‰,

along with a decrease of lower magnitude in the tropical sea surface temperature.

Although this mechanism is view as a long term process (Gregory, 1991 in Shields et al.,

2003) and thus failing to elucidate the short term of the δ18O SMOW change, more 164

Figure 11.4: Temperature vs. calcium carbonate shell δ18O (PDB) for various values of seawater δ18O (SMOW) assuming an ice-cap free world. Even maximum δ18O values for the Early Ordovician and early Middle Ordovician (-5.4‰and -6.3‰, respectively) yield unrealistic minimum seawater temperatures (37°C – 43°C) if δ18O seawater was -1‰ SMOW, while δ18O seawater of -3‰ SMOW yields more plausible minimum tropical sea-surface temperatures of 27°C – 32°C (Shields et al., 2003).

recent modeling studies indicate that such δ18O SMOW changes can occur over much shorter periods of time (5 to 50 My), particularly during periods of high oceanic spreading rates (Lécuyer and Allemand, 1999). The short-term δ18O SMOW shift theory is consistent with the inferred high oceanic spreading rates supposed to characterize the

Caradocian based on the 87Sr/86Sr isotopic ratio (Qing et al., 1998; Shields et al., 2003).

In a world characterized by energetic intra-basinal activities, that would generate not only higher input levels of nutrients (favoring high organism productivity), but also 165 abundant CO2, a widespread anoxia event can occur. The more anoxic the bottom waters become, the more the vertical partitioning of the oxygen isotopes takes place, generating positive shifts in the δ18O. Consequently, a sharp positive shift in the δ18O documented at the beginning of the Caradocian, could also reflect a strongly stratified water column

(Tobin et al., 1996), with the heavier oxygen entrapped in the bottom-most water layers, and not necessarily a climatic cooling. The energetic volcanic activity in the southern margin of Laurentia is also supported by two regionally widespread Middle Caradocian air-fall ash beds, the Deicke and Millbrig k-bentonites (Huff and Kolata, 1990).

11.3.2. The biotic system

Caradocian biotas witnessed high taxonomic diversity for some groups of benthic and planktonic organisms (e.g., radiolarians, graptoloids, brachiopods; fig. 11.1A), and a decreased diversity and extinction for other groups (Sloan, 1988a; 1988b; Patzkowsky and Holland, 1993). High organism productivity is shown by a middle Caradocian positive excursion in the carbon isotopic values (Patzkowsky et al., 1997; Pancost et al.,

1999).

Sea level during this epoch was high (Ross and Ross, 1992; Qing et al., 1998), so that the Caradocian chert peak and organism radiation coincide with an important transgression (fig. 11.1B). The Caradocian transgression is believed to be the largest in the Phanerozoic (Ross and Ross, 1992) due to fast oceanic floor spreading resulting in high ridge volume (Qing et al., 1998). Although extinction events have been linked to transgressions through the invasion of the shelf area by anoxic waters (Wilde and Berry,

1984), in this instance there was a proliferation of a variety of benthic and planktonic 166 groups: radiolarians (Vishnevskaya, 1997), graptolites, brachiopds (Leggett et al., 1981)

(fig. 11.1A), and siliceous sponges (Rigby and Webby, 1988). Higher diversities and revolutions in the marine realm have been correlated on one hand with cooling episodes and low nutrient supply (Lipps, 1970), and on the other hand with high submarine volcanism and hydrothermal activity providing a thermal stimulus (Hays and Pittman,

1973; Vermeij, 1995).

The Caradocian radiation was concentrated around regions of active mountain building and volcanisms (e.g., the Iapetus Ocean), which suggests a direct relationship between organism diversity and the proximity of tectonic activity (Vermeij, 1995; Miller and Mao, 1995; Miller, 1997). In addition, many Caradocian deep-sea chert deposits formed in active areas such as the southeastern part of Laurentia characterized by arc- continent collision tectonism, and Kazakhstan, which was a massive island-arc volcanic region (Nikitin et al., 1991), supporting the high diversity - high productivity - high volcanism inter-relationship model.

11.3.3. Considerations on the chert abundance

Slope/basinal environments. Why a chert peak?

The bottom water environment: anoxia?

At the bottom of basins, it appears that anoxic waters develop as a consequence of the converging action of two factors: high CO2 input due to high volcanic – mid-oceanic activity, and high input of organic matter from the surface layers due to an increased productivity volcanically induced (fig. 11.5). Oxygen-depleted bottom water model 167

would be consistent with the observed abundance of graptolite-rich black shale, (Berry

and Wilde, 1978; Leggett et al., 1981), abundant volcanogenic massive sulfide deposits

(VMS) identified in the Iapetus Ocean and marginal basins (Eastoe and Gustin, 1996), and abundant oolitic ironstone accumulations (Van Houten and Arthur, 1989) (fig.

11.1A&B). However, whereas the Caradocian age is well documented for the black shale

peak (Leggett et al., 1981), the timing of the VMS and oolitic ironstones abundance

peaks is less well established.

The analyses of Phanerozoic massive sulfide deposits (Eastoe and Gustin, 1996) and oolitic ironstones (van Houten and Arthur, 1989) and their associated lithologies,

emphasized their frequent association with black shales. Sulfide minerals, as well as

black shales, may be preserved either in oxygen depleted environments (the preservation

model), or due to high organic productivity accompanied by rapid burial under well

aerated conditions (the productivity model) (Calvert, 1987; Morris, 1987; Arthur and

Sageman, 1994; Eastoe and Gustin, 1996). However, considering that modern VMS

deposits that form in oxygenated water on the mid-oceanic ridges are preserved to a small extent, it has been inferred that large accumulations of VMS are tightly linked to anoxic bottom waters (Eastoe and Gustin, 1996).

Based on the common association between massive sufide deposits, oolitic ironstones, and black shales, along with the Caradocian age of the Ordovician black shale peak (Leggett et al., 1981), it can be speculated that the VMS peak along with that of the oolitic ironstones may be of Caradocian age as well. According to Eastoe and Gustin 168

Figure 11.5: Chert abundance Model. CARADOCIAN [458 – 449 Ma] 169

(1996), the abundance of VMS in the Iapetus Ocean and proximal basins during the

Ordovician, is the result of the interplay between anoxic waters and tectonic factors.

Caradocian VMS and oolitic ironstone abundances would be consistent with the

Caradocian climax of the mid-oceanic ridge activity inferred from the strontium isotopic

ratio (Shields et al., 2003) that, in turn, generated a high sea level, conditions that

represent the most favorable ambient for widespread anoxia (Mackenzie and Piggott,

1981). Documenting extensive black shale deposits for the Ordovician and Silurian,

Fischer and Arthur (1977) remarked on their association with transgression phases and

predicted the Caradocian chert peak based on the inference that during a polytaxic

interval (a greenhouse oxygen-depleted state; Fischer and Arthur, 1977), deep water

sediments should be characterized by more widespread cherts than those that form during

olygotaxic times.

The sole presence of black shales, or massive sulfide deposits, does not indicate

anoxic waters, however their coexistence along with oolitic ironstones in a world

characterized by energetic intra-basinal activity, such as high mid-oceanic spreading

rates, marine volcanism and hydrothermal activity, point toward an euxinic environment.

Within such an environment opal accumulation and burial would be greatly enhanced

(e.g., De Wever et al., 1994).

The surface waters: low or high organism productivity?

A variety of evidence points toward an anoxic deep water environment during the

Caradocian, but whether or not it was accompanied by a high productivity is unclear. It is known that radiolarians, graptoloids, and brachiopods record high taxonomic diversities 170

during the Caradocian (fig. 11.1A). Even though a high diversity and abundance

commonly show a reverse correlation (Tappan and Loeblich, 1973), studies of modern

radiolarian faunas from the East China Sea show that high productivity and high diversity can co-occur (Chang et al., 2003).

The high productivity and organic carbon burial model proposed for the surface water planktonic organism activity is in agreement with a middle Caradocian (P. tenuis –

P. undatus conodont zones) positive excursion in the carbon isotope composition of carbonate (~ 3‰) and of organic carbon (~ 4‰-6‰) (Patzkowsky et al., 1997; Pancost et al., 1999). Huff and Kolata (1990) identified two regionally widespread air-fall ash beds, the Deicke and Millbrig k-bentonites, of the exact same age (P. tenuis – P. undatus

interval), which is viewed as chronostratigraphic marker horizons throughout much of the

eastern Mid-Continent. The question is did this volcanic pulse induce indirectly the

carbon spike by killing organisms? The carbon spike has been also correlated with

Caradocian sequences in Estonia (Ainsaar et al., 1999, in Shields et al, 2003), suggesting

an event of larger magnitude than one confined to the eastern margin of Laurentia. If this

volcanic burst affected marine biotas to such a wide extent as to imprint a strong carbon

shift in the rock record, we can ask if it was not also responsible for a short-lived

warming driven by the high amount of CO2 exhaled into the atmosphere. The oxygen

data of Shields et al. (2003) show a sharp, short-lived negative shift during the same age interval (fig.11.1B) also coeval with a sea level rise, that could be the result of the same large-scale volcanic event. All these converging phenomena provide evidence in support of the greenhouse hypothesis, at least for the early and middle Caradocian. Further, a 171

warm Caradocian model implies that the cooling mode did not start until the late

Caradocian, which means that the middle Caradocian carbon shift does not reflect a high

organism productivity. However, it is possible that the cooling mode could have been triggered by rapid CO2 drawdown resulting from high biotic productivity. More data are

needed in order to test a Caradocian greenhouse mode based on the above evidence.

The Caradocian chert peak is coeval with high oceanic spreading rates, energetic

volcanism, and hydrothermal activities that are consistent with a high sea level (Qing et al., 1998). Vigorous intra-basinal volcanic activity would stimulate higher silica and other

nutrient fluxes so that the productivity of radiolarians and other planktonic organisms

would be boosted inasmuch as the surface water temperatures are still favorable.

Energetic intra-basinal activities may initiate energetic vertical water advection, including

coastal upwelling (volcanic-induced upwelling of Vogt, 1989). Thus, increased nutrient

input has two sources: mid-oceanic activity, and upwelling of rich-nutrient sediments.

Numerous studies have stressed the linkage between high biogenic chert abundance, high

silica secreting organism productivity, and intensive volcanic/tectonic/magmatic activity

(Grunau, 1965; Steinberg, 1981; Aubouin, 1965; Racki, 1999; Racki and Cordey, 2000).

If we accept the initiation of a climatic cooling at the beginning of the

Caradocian, as suggested by the oxygen isotopic ratios, and taking into account that the

solubility of silicic acid in water decreases with temperature, in a cooler environment less

silica is dissolved in the water column and thus more opal would reach the seafloor to be

preserved as chert. A cooling would affect principally the surface waters in contrast to

much more stable deep waters. Modern studies show that 50% of the total biogenic silica 172

production is remobilized in the upper layers of the ocean (the upper 100 m) and

therefore a coling would increase the potential of opal accumulation and burial through

an increased opal rain rate. Also, silica dissolution depends directly on water temperature, most of the opal accumulation occurring in areas where the surface water is cold

(DeMaster, 1981). Because large temperature variations occur within the surface layers of water (100-200 m) and the deep-sea is confined to a relatively narrow range of low temperatures (Bidle et al., 2002), a very important factor in preservation of opal in basinal settings is ultimately the temperature of surface layers of water (Nelson et al.,

1995). However, a decrease in the opal dissolution rate in the surface waters would be triggered by lower water temperature. The magnitude of a potential Caradocian seawater cooling is constrained by the high diversity of radiolarians whose metabolism in modern

settings is negatively affected by low temperatures (Anderson et al., 1989c).

In conclusion, higher opal accumulation and burial rates would be generated by:

1) a global productivity acme in coastal and abyssal settings triggered by high

endogenous inputs of silica and other nutrients; and 2) excellent taphonomic conditions

within an anoxic environment driven by the high fluxes of organic matter, CO2 and other oxygen-consuming cations. This model suggests that preservation and productivity do not exclude each other as mechanisms that induce the formation of cherts, and that their concurring action only generates a higher chert peak, under the incidence of tectono- volcanically induced taphonomic, oceanographic, and biosiliceous productivity conditions (fig. 11.5). A surface water cooling could add even more to the potential for 173

opal accumulation due to lower dissolution rates under the circumstances the surface

water temperature still represents a favorable environment for radiolarians.

Shelf settings

During the Caradocian, widespread carbonate platforms, representing a good

environment for the development of siliceous sponge faunas, were again dominating the

shelf system. Siliceous sponges that at this point were inhabiting the deeper sites of the

shelf area, could have taken advantage of a volcanically-induced coastal upwelling of the

nutrient rich waters. Strontium data show strong non-radiogenic values indicating a

nutrient input cycle dominated by endogenous fluxes and a possible seawater cooling

would not affect too strongly this opportunistic group capable of inhabiting shallow to

deep environments since the Early Cambrian (Finks, 1960), and low to high latitudes at

least since the Lower Ordovician (Beresi and Esteban, 2004). After a period of ecological adaptation to new ecological niches after their migration from onshore toward more offshore settings during the Middle Ordovician, siliceous sponges would proliferate in a stable, nutrient rich carbonate environment. Chert burial would be increased due to higher silica levels via siliceous sponges, and possibly due to a lower opal dissolution rate in cooler seawater.

11.4 Upper Ordovician: the Ashgillian - an icehouse period

During the Ashgillian, the chert abundance decreases in all three depositional environments (Table 7.2.B and 7.3.B), but it is still higher than the Middle Ordovician chert low. While the formation of peritidal/lagoonal chert deposits ceases, shelf cherts 174 dominate by a factor of two over the slope/basinal deposits. The slope/basinal chert deposits exhibit the first drop in abundance for the Ordovician. The shelf cherts that dominate the Ashgillian biosiliceous burial surpass both the Lower and the Middle

Ordovician shelf chert occurrences. The distribution of silica-secreting organisms within basins show no changes - planktonic radiolarians are still confined to slope/basinal settings, and siliceous sponges to shallow and deep water environments (shelf and slope/basin). Taking into account the genetic relationship that exists between radiolarians and slope/basinal bedded chert formation (Chapter 7 and 9), it seems that radiolarian fauna played an important part in this general chert abundance decrease. From the point of view of the geographical distribution of cherts, both Laurentia and Gondwana witness significant decreases in silica burial based on the decrease in the number of chert deposits

(Table 7.6).

11.4.1. The abiotic system

The Ashgillian was characterized by the presence of ice sheets (over both poles,

Poussart et al., 1999), low sea levels, and the ongoing Taconic orogeny. Considerable land area was exposed by the glacially-induced sea level drop whose magnitude has been estimated between 50m and 100m (Brenchley and Newall, 1980; Finney, 1997), and by the ongoing Taconic orogeny, two physical processes that induce the exposure of fresh silicate rocks. However, the strontium data still exhibit strong non-radiogenic values and no differences between the Caradocian and the Asghillian (fig. 11.1B; Shields et al.,

2003). These values have been explained by a delay in the continental strontium influx 175

caused by the time necessary for erosion to remove the emerged widespread carbonate

shelves before exposing the more radiogenic rocks underneath (Qing et al.,

1998). The oceanic circulation underwent an overturn through the formation of ice sheets

that cause the sinking of denser oxygenated waters at the poles, thus inducing the

formation of active coastal upwelling and an increase in wind, wave and storm action as a

consequence of larger thermal gradients between the poles and the tropics (Berry and

Wilde., 1978).

Short-lived oxygen (3‰-4‰) and carbon (as much as 7‰) excursions have been

used as arguments to confine the timing of the glaciation to the late Ashgillian, the

Hirnantian stage, suggesting a short-lived glaciation (Brenchley et al., 1994, 2003; Gibbs et al. 1997). Muehlenbachs (1998) asserts that during glacial periods the seawater δ18O can increase with up to 2‰ - 3‰ assuming an originally ice-free world. Mechanisms that have been proposed to explain such a short-lived glaciation include high organism productivity (Brenchley et al., 1994) and the Taconic orogeny (Kump et al., 1999), which may have initiated cooling followed by a more dramatic shift into full icehouse conditions. Global biochemical models suggest that the Late Ordovician glaciation has occurred under high pCO2, levels between 8 and 16 times higher than today’s level

(Berner, 1994; Kump et al., 1999; Yapp and Poths, 1992; Gibbs et al., 1997; Herrman et al., 2003).

11.4.2. The biotic system

The Late Ordovician extinction, considered one of the five big Phanerozoic extinctions (Sepkoski, 1995), took place in two phases (two major faunal turnovers) 176

related to two climatic and environmental changes: first, a cooling period ending with the onset of the glaciation, secondly, a deglaciation and warming event (Brenchley, 1989;

Brenchley et al., 2003 the second phase characterized by deglaciation and warming witnessed most part of the late Ordovician extinction (Brenchley et al., 2003).

Whether or not the late Ordovician glaciation affected radiolarian faunas is controversial. Noble et al. (1997) reported a lack of major radiolarian extinction while

Vishnevskaya (2000) noted a major radiolarian extinction across the Ordovician-Silurian boundary (Ashgillian – Llandoverian) with radiolarian diversity droping from 100 to 20 species (fig. 11.1A). Following this drop, the rest of the Silurian witnesses a slight increase in diversity upwards to 60 species (Vishnevskaya, 2000). However, to date there are no known individual rock sequences in which extinction in the siliceous radiolarian

fauna is clearly recorded.

Relative to the Ordovician silicisponges, within a shallowing interval of the lower

Hanson Creek Formation (late Upper Ordovician) Finney et al. (1999) recognize a faunal

turnover in siliceous sponges, defined by a change from deep water, hexactinellid-

dominated assemblages, to shallow, subtidal lithistid-dominated assemblages as a

consequence of the sea level drop.

11.4.3. Considerations on the chert abundance (fig. 11.6)

During glacial periods abundant chert accumulation should occur at upwelling

sites due to a high organism productivity. Within the shallower facies the amount of

siliceous deposits could decrease as a consequence of changes in the ecological

distribution of siliceous organisms driven by the glacial-eustatic sea level regression. 177

Figure 11.6: Chert abundance Model. ASHGILLIAN [449 – 443 Ma 178

Moreover, a drop in sea level would increase the volume of silicates available to

weathering, as well as the dissolved silica input into the basinal facies, while the shelf is

more or less emerged. The amount of dissolved silica in seawater could be higher and

more offshore chert may occur.

Shelf environments

Shelf water environments were populated exclusively by the silicisponges during

the Ashgillian. Even though the shelf chert abundance registers a decrease as compared to

the Caradocian, the normalized data show that Ashgillian shelf occurrences rank second

in abundance even under the circumstances of a severe sea level drop (50-100m) and

exposure of shelves. This finding suggests two main posibilities:

1) The siliceous sponge faunas were not affected significantly by seawater

cooling induced by the Late Ordovician glaciation. This view is supported by the ecology of modern silicisponges which demonstrates that they tolerate large thermal ranges, paleolatitudinally as well as vertically within the basin (Reincke and Barthel, 1997). The

low to mid-paleolatitudinal distribution of silicisponges shown by the present study

(Chapter 10), also supports the interpretation that siliceous sponge faunas were not

significantly affected by the cooling.

2) The glacial-eustatic sea level drop was of short duration so that the induced

loss of living habitats did not affect the opal production and burial to any great extent.

The normalized data demonstrates a high shelf chert abundance during the Ashigillian,

outnumbered only by the Caradocian siliciponge-induced shelf cherts. Moreover, silicisponges are able to inhabit varied ecological niches, carbonate or not, as long as they 179

have clear waters and sufficient nutrients. Under these circumstances, the fact that they

were populating the shelf environment during the Ashgillian demonstrates low

terrigenous inputs, contrary to the high input that we would expect as a consequence of

the continental weathering of fresh silicate rocks exposed by the Taconic orogeny. This

hypothesis is consistent with the strontium ratio values that still show non-radiogenic

characteristics. At the other and of the spectrum, a long-term sea level drop would

drastically affect siliceous sponge habitats and implicitly the silica burial. An analysis

limited just to Ashgillian siliceous sponge shelf cherts would not offer insight into their

opal production potential and the factors that influence it. However, such an analysis is

clearly put in perspective when performed in the context of the whole Ordovician.

Slope/basinal environments

Because this survey does not differentiate between the slope and deep sea

(basinal) environments of chert formation, two oceanographic systems that function differently with respect to opal accumulation and preservation (Nelson et al., 1995), I can

only speculate about the implications of the glacially-induced effects on biosiliceous

productivity, opal accumulation and burial. The impact of a vigorous oceanic circulation

through vertical advection of cold, nutrient-rich, deep waters, might generate volumetrically important chert deposits spatially concentrated at upwelling sites rather than widespread abundant smaller cherts (Kidder, oral communication). In view of this 180

fact, estimation of chert abundance based on frequency of occurrences becomes futile in

interpreting any change in silica burial caused by glacially induced oceanic turnovers.

High chert abundances, on the slope as a result of high organism productivity induced by upwelling, would imply that the silica secreting biota responsible for chert formation behave as well in cooler environments as in warm waters. During the

Ashgillian, radiolarians were very likely inhabiting the whole transect of the deep part of a basin, from slope to deep-sea, and thus they were under direct influence of the coastal upwelling. The oxygen isotopic ratios show an important cooling during this interval (11

- 17°C; Shields et al., 2003). To date both modern data (Sugyiama and Anderson, 1997;

Anderson et al., 1989, a,b,c) and paleo-ecological results (the present study) show the affinity of radiolarians with warm environments, demonstrating that environmental factors such as temperature and salinity are more significant limiting factors in radiolarian growth and longevity.

The slope/basinal bedded chert drop therefore is interpreted as a negative perturbation of radiolarian productivity due to cold surface seawater. In deep-sea settings, within a scenario of cooler oceanic surface waters, the lower silica dissolution rate would increase the potential for higher preservation and chert formation. However, such a high preservation is not substantiated in the rock record, which supports the hypothesis that the slope/basinal bedded chert drop was a result of productivity slowdown due to low water temperature. Moreover, a general high organism productivity, induced by the oceanic overturn, associated with the whole Ashgillian interval has not been demonstrated either by abundant black shale deposits (Leggett et al., 1981), or by positive excursions in δ13C. 181

The first important Ashgillian change in the amount of organic carbon burial has been

documented as occurring in the Hirnantian, the youngest time-defined level of the

Ashgillian (Brenchley et al., 1994).

It is worth mentioning that the Ashgillian shelf and slope/basinal cherts surpass

and almost equal those of the Lower Ordovician in spite of the sharp decline in silica burial recorded between the Caradocian and the Ashgillian. Silica output into cherts expresses ultimately the balance between biogenic silica production, dissolution, dilution by host sediment, and preservation rates. Therefore two different mechanisms can be considered responsible for triggering the chert formation corresponding to the two antagonistic climatic modes that characterize the Lower Ordovician and the Ashgillian: high biosiliceous productivity, associated with good taphonomic bottom water conditions within a greenhouse mode (Lower Ordovician) vs. low biosiliceous productivity associated with low dissolution rate within an icehouse mode (Ashgillian). 182

Chapter 12

CONCLUSIONS

A global survey of the spatial and temporal distribution of Ordovician nodular and bedded cherts, and subordinately of silicified oolites, was undertaken to address three questions: 1) Did the Early Paleozoic radiolarians and siliceous sponges behave similarly to diatoms regarding their response to abrupt changes in oceanic circulation and climate parameters? 2) Did cherty facies retreat from peritidal settings during the Ordovician? 3)

Can chert deposits be of use in evaluating whether the late Ordovician glaciation spanned most of the Caradocian and Ashgillian (Pope and Steffen, 2003) or was the glaciation more tightly constrained to within the Ashigillian (Brenchley et al., 1994)? Further on, the volume and resolution of the data allowed quantitative and qualitative analyses of the relationships between chert deposit types, depositional environments, and silica secreting organisms.

The results of this study can be summarized as follows:

1) Most Ordovician chert deposits are of biogenic origin due to: a) approximately

38% of the Ordovician chert formations contain siliceous skeletal remains; and b) the onshore-offshore migration of the silica secreting organisms, perceived through the rock record of chert bearing siliceous organisms, is accompanied by an onshore-offshore change in the locus of silica burial, reflected in the stratigraphic record of cherts that lack

preserved siliceous skeletal remains. 183

2) The analysis of chert types and reported skeletal remains within cherts reveals a segregation of chert types due to a genetic relationship between the silica secreting organisms and the types of the chert deposits. Peritidal/lagoonal and shelf cherts are mainly nodular cherts hosted by carbonate facies. These cherts appear to have formed predominantly via silica secretion by lithistid demosponges. Most of these cherts are low latitude occurrences (< 30ºN and S). Slope/basinal cherts are predominantly bedded cherts, mainly hosted by fine-grained siliciclastic sediments, and bear radiolarians. Most of these cherts are also confined to low latitudes. At mid latitudes, the skeletal components of these cherts are mainly hyalosponges. Also, the analysis showed that the formation of some shelf bedded cherts might by induced by silicisponges.

3) The Middle Ordovician distribution pattern of silica secreting biotas reflects a migration of siliceous sponge faunas from peritidal/lagoonal to shelf-basinal environments, and of radiolarians to basinal settings, suggesting an Arenigian-

Llanvirnian onshore-offshore migration of the silica secreting organisms. This refines the timing of an earlier suggested retreat in the locus of silica burial during this period. It also indicates that Early Paleozoic planktonic faunas accompanied the benthic communities.

Whether the timing of the retreat matches the roughly coeval onshore-offshore shift of the benthic Cambrian and Paleozoic faunas (Jablonski et al., 1983) is unclear. The Middle

Ordovician minimum of peritidal/lagoonal and shelf cherts may be a result of shelf area reduction during a relatively low sea level. However, the reappearance of shelf cherts in the Caradocian indicates that, although silicisponges did retreat significantly from peritidal/lagoonal settings in the early Middle Ordovician, they still largely populated the 184

shelf settings. The few Middle Ordovician and Caradocian silicified oolites suggest that

silicisponges may have still been present sporadically in some peritidal/lagoonal settings

that were characterized by peculiar local conditions.

4) The Ordovician spatial and temporal distribution patterns of siliciponges and

radiolarians suggest that silica limitation governed the distribution of silicisponge faunas,

and that physical factors such as temperature played a major role in the distribution of

radiolarians.

5) Lower Ordovician cherts are abundant and widespread, occurring in equal

proportions in peritidal/lagoonal, shelf, and slope/basinal settings. Within a warm world

characterized by stable and favorable environmental conditions, high sea level, and

widespread carbonate platforms, the chert abundance in peritidal/lagoonal and shelf

environments seems to be the result of the interplay between good organism productivity

and enhanced chert preservation. The dominance of bedded cherts in slope/basinal

settings may reflect relatively high biosiliceous productivity, probably associated with

favorable taphonomic conditions under an oxic regime within the bottom waters.

6) A significant decrease in the number of chert deposits occurs in the Middle

Ordovician. While the peritidal/lagoonal and shelf environments record a strong decline

in silica burial, the slope/basinal bedded cherts record a slight augmentation. This finding suggests that the redistribution of benthic shallow water siliciponges played an important

role in this general chert abundance low. The integration of data on biotic and abiotic

changes that occurred during the Middle Ordovician indicates that the Middle Ordovician

chert low was driven by a shortage in the shallow water silica burial due to redistribution 185

of silicisponges and their habitat and subsequently to a lesser extent of radiolarians,

possibly accentuated by a general Middle Ordovician sea level low. The slightly

increased slope/basinal chert abundance during this interval seems to conform to a good

organism productivity, enhanced preservation rates and aerated bottom waters model.

7) A resurgence in the number of chert deposits occurs in the Upper Ordovician, more precisely during the Caradocian. This occurs in both shelf and slope/basinal environments, where chert abundance surpasses that of the Lower Ordovician by 50%.

However, Upper Ordovician chert abundance in peritidal/lagoonal settings remained at essentially the same low levels as in the Middle Ordovician.

8) The Caradocian marks the highest point in the abundance of Ordovician cherts, that appear to have accumulated during a warming interval. This chert peak is coeval with a transgression, high abundance of black shales, oolitic ironstones, and volcanogenic massive sulfides, as well as the Ordovician peaks in diversity of radiolarians and other organisms, are all consistent with a warm stable world and / or oceanic anoxia. Changes in δ18O values at the Middle-Upper Ordovician boundary are consistent with warming

late in the Caradocian. Values of 87Sr/86Sr shift toward non-radiogenic values and the sea

level toward transgression. These suggests that continental input of dissolved silica

declined instead of rising in the Caradocian, despite the coeval Taconic orogeny. The

high chert abundance may have been the result of a global productivity acme triggered by

high endogenous inputs of silica and other nutrients, along with excellent taphonomic

conditions within an anoxic environment driven by the high fluxes of organic matter,

CO2, and other oxygen-consuming cations. 186

9) The decline of cherts during the Ashgillian, when the glaciation was clearly in effect, was not expected based on assumptions of modern silica accumulation in an icehouse with vigorous ocean circulation. While the slope/basinal cherts record a strong decline, shelf cherts exhibit only a fairly low reduction. This finding suggests that radiolarians played an important role in this relatively low general chert abundance. The decline may reflect: a) a negative perturbation of radiolarian activity due to cold surface seawater; b) accumulation of fewer but volumetrically more important basinal cherts reflecting upwelling settings in the Ashgillian, under the circumstances where radiolarians behave as well as in cold waters as in warm waters; c) depletion of oceanic dissolved silica content via silica burial during glacially induced intensive upwelling again under the circumstances where low temperatures are not a limiting factor for radiolarian growth. The relatively high silicisponge-induced shelf chert abundance within an icehouse world outnumbered only by the Caradocian shelf cherts could be linked to a glacially-induced sea level drop of short duration. An analysis limited just to Ashgillian silica secreting organisms would not offer insight into their behavior and opal production potential and the factors that influence them. However, such an analysis is clearly put in perspective when performed in the context of the whole Ordovician, by epoch. Chert abundance alone probably will not resolve the debate about the duration of the

Ordovician glaciation. However, consideration of their abundance patterns in conjunction with other aspects of the Earth system is consistent with Caradocian warming followed by a brief Ashgillian cooling and glaciation, the causes of which remain enigmatic. 187

REFERENCES:

Abelmann, A. and Gowing, M.M., 1997. Spatial distribution pattern of living polycystine radiolarian taxa - baseline study for paleoenvironment reconstructions in the (Atlantic sector). Marine Micropaleontology, 30(1-3): 3-28.

Agnew, A.F., 1955. Facies of Middle and Upper Ordovician rocks of Iowa. AAPG Bulletin, 39: 1703-1752.

Ainsaar, L., Martma, T. and Meidla, T., 1999. Evidence for a widespread carbon isotopic event associated with late Middle Ordovician sedimentological and faunal changes in Estonia. Geological Magazine, 136: 49-62.

Aitchison, J.C., 1998. A Lower Ordovician (Arenig) radiolarian fauna from the Ballantrae Complex, Scotland. Scottish Journal of Geology, 34(1): 73-81.

Aitchison, J.C., Flood, P.G. and Malpas, J., 1998. Lowermost Ordovician (basal Tremadoc) radiolarians from the Little Port Complex, western Newfoundland. Geological Magazine, 135(3): 413-419.

Anderson, O.R., Bennett, P., Angel, D. and Bryan, M., 1989a. Experimental and observational studies of rariolarian physiological ecology: 2. Trophic activity and symbiont primary productivity of Spongaster tetras tetras with comparative data on predatory activity of some Nassellarida. Marine Micropaleontology, 14: 267- 273.

Anderson, O.R., Bennett, P. and Bryan, M., 1989b. Experimental and observational studies of rariolarian physiological ecology: 3. Effects of temperature, salinity and light intensity on the growth and survival of Spongaster tetras tetras mainained in laboratory culture. Marine Micropaleontology, 14: 275-282.

Anderson, O.R., Bennett, P. and Bryan, M., 1989c. Experimental and observational studies of rariolarian physiological ecology: 1. Growth, abundance and opal productivity of the spongiose radiolarian Spongaster tetras tetras. Marine Micropaleontology, 14: 257-265.

Anderson, O.R., Bryan, M. and Bennett, P., 1990. Experimental and observational studies of rariolarian physiological ecology: 4. Factors determining the distribution and survival of Didymocyrtis tetrathalamus tetrathalamus with implications for paleoecological interpretations. Marine Micropaleontology, 16: 155-167.

188

Andrichuk, J.M., 1959. Ordovician and Silurian stratigraphy and sedimentation in southern Manitoba. AAPG Bulletin, 43: 2333-2398.

Armstrong, H., Owen, A. and Floyd, J., 1999. Rare earth geochemistry of Arenig cherts from the Ballantrae Ophiolite and Leadhills Imbricate Zone, southern Scotland: implications for origin and significance to the Caledonian Orogeny. Journal of the Geological Society, 156(3): 549-560.

Armstrong, H., Rushton, A., Owen, A. and Floyd, J., 1998. Biostratigraphy of the Currarie Formation on the northwestern edge of the Southern Uplands: Implications for the cessation of basic volcanism. Scottish Journal of Geology, 34(2): 119-125.

Arthur, M.A. and Sageman, B.B., 1994. Marine black shales: depositional mechanisms and environments of ancient deposits. Annual Reviews Earth and Planetary Sciences, 22: 499-551.

Astini, R.A., 1995. Geologic meaning of Arenig-llanvirn diachronous black shales (Gualcamayo Alloformation) in the Argentine Precordillera, tectonic or eustatic? In: J.D. Cooper, M.L. Droser and S.C. Finney (Editors), Ordovician Odyssey : short papers for the Seventh International Symposium on the Ordovician System. The Pacific Section Society for Sedimentary Geplogy, pp. 217-220.

Aubouin, J., 1965. Geosynclines. Elsevier Pub. Co., Amsterdam, 335 pp.

Barron, J.A. and Baldauf, J.G., 1995. Cenozoic marine diatom biostratigraphy and applications to and paleoceanography. In: E.B.B.W.I. Williams (Editor), Siliceous - Short courses in paleontology. The Paleontological Society, pp. 107-118.

Bauert, H., 1994. Petrology and geochemistry of Middle Ordovician foredeep and platform carbonates at two locations in the Appalachian Basin. Thesis (Master's) Thesis, University of North Carolina, Chapel Hill, 102 pp.

Bengtson, S., 1986. Siliceous microfossils from the Upper Cambrian of Queensland. Alcheringa, 1013-4(195-216).

Bengtson, S., 1994. The advent of animal skeletons. In: S. Bengtson (Editor), Early life on Earth. Nobel Symposium. Columbia University Press, New York, pp. 412-425.

Bennison, A.P., 1975. Geological Highway Map of the Southeastern Region Alabama, Georgia, Florida, Louisiana, Mississippi. AAPG, Tulsa, Oklahoma.

189

Bennison, A.P., 1989. Geological Highway map of the Mid-Atlantic region. AAPG.

Beresi, M. and Rigby, J.K., 1993. The Lower Ordovician sponges of San Juan, Argentina. Brigham Young University Geology Studies., 39: 1-64.

Beresi, M.S. and Esteban, S.B., 2004. Sponge Fauna and Spicule Assemblages from the Ordovician of Argentina: A Review. Instituto Superior De Correlación Geológica (INSUGEO). Serie de Correlación Geológica. Aspects of the Ordovician System in Argentina, 16: http://www.unt.edu.ar/fcsnat/INSUGEO/correlacion16/libro_16.htm.

Berner, R.A., 1994. 3GEOCARB II: a revised model of atmospheric CO2 over Phanerozoic time. American Journal of Science, 294: 56-91.

Berry, W.B.N., 1958. Depositional environment of the Ordovician succession in the Marathon region, Texas. Journal of Sedimentary Petrology, 28(4): 389-405.

Berry, W.B.N., 1962. Comparison of Some Ordovician Limestones. AAPG Bulletin, 46(9): 1701-1720.

Berry, W.B.N., Quinby-Hunt, M.S. and Wilde, P., 1995. Impact of Late Ordovician glaciation-deglaciation on marine life. In: C.o.G. Board on Earth Sciences and Resources, Environmaen, and Resources, National Research Council (Editor), Effects of past global change on life. National Academy Press, Washington, D.C., pp. 34-46.

Berry, W.B.N. and Wilde, P., 1978. Progressive ventilation of the oceans - an explanation for the distribution of the Lower Paleozoic black shales. American Journal of Science, 278: 257-275.

Beus, S.S., 1968. Paleozoic Stratigraphy of Samaria Mountain, Idaho-Utah. AAPG Bulletin, 52(5): 782-808.

Bidle, K. and Azam, F., 1999. Accelerated dissolution of diatom silica by marine bacterial assemblage. Nature, 397(6719): 508-512.

Bidle, K., Manganelli, M. and Azam, F., 2002. Regulation of oceanic silicon and carbon preservation by temperature control on . Science, 298(5600): 1980-1984.

Bird, J.M., 1989. Stratigraphy and structure of the northern part of the Fish Creek Range, Eureka County, Nevada. Master's Thesis, Oregon State University, Corvallis, 143 pp.

190

Blome, C.D., Hull, d.M., Pessagno, E.A., Jr and Reed, K.m., 1995a. Mesozoic Radiolaria. In: C.D. Blome, P.M. Whalen and K.M. Reed (Editors), Siliceous microfossils. Short courses in paleontology ; no. 8. [Knoxville, Tenn.] : Paleontological Society, pp. 31-60.

Blome, C.D., Whalen, P.M. and Reed, K.M., 1995b. Siliceous microfossils, 8. Paleontological Society, 185 pp.

Boggs, S.J., 1987. Principles of sedimentology and stratigraphy. Merrill Publishing Company, Columbus, Ohio, 784 pp.

Borchiellini, C., Chombard, C., Lafay, B. and Boury-Esnault, N., 2000. Molecular systematics of sponges (Porifera). Hydrobiologia, 420: 15-27.

Borchiellini, C., Manuel M., Alivon, E., Boury-Esnault, N., Vacelet, J., Le Parco, Y., 2001. Sponge and the origin of Metazoa. Journal of Evolutionary Biology, 14(1): 171-179.

Boury-Esnault, N. and Vacelet, J., 1994. Preliminary studies on the organization and development of a hexactinellid sponge from a Mediterranean cave, Oopsacas minuta. In: R.W.M. van Soest, van Kempen, Theo M. G., Braekman, Jean- Claude, Brink, Anita, Bianchi, Frans R., Vermuelen, Jan J. (Editors), Sponges in time and space. Biology, Chemistry, Paleontology. A. A. Balkema, Rotterdam, pp. 407-415.

Bramlette, M.N., 1946. The Monterey Formation of California and the origin of its siliceous rocks. U.S. Geol. Survey Professional Papers, 212: 55.

Brasier, M., Green, O. and Shields, G., 1997. Ediacarian sponge spicule clusters from southwestern Mongolia and the origins of the Cambrian fauna. Geology, 25(4): 303-306.

Brasier, M.D., 1992. Paleoocenography and changes in the biological cycling of phosphorus across the Precambrian-Cambrian boundary. In: J.H. Lipps and P.W. Signor (Editors), Origin and earlyevolution of the Metazoa. Topics in . Plenum Press, New York, pp. 483-523.

Brenchley, P.J., 1989. The Late Ordovician extinction. In: S.K. Donovan (Editor), Mass extinctions: processes and evidence. Columbia University Press, New York, pp. 104-132.

191

Brenchley, P.J., Carden, G. A., , Hints, L., Kaljo, D, Marshall, J. D., Martma, T., Meidla, T., Nolvak, J., 2003. High-resolution stable isotope stratigraphy of Upper Ordovician sequences; constraints on the timing of bioevents and environmental changes associated with mass extinction and glaciation. Geological Society of America Bulletin, 115(1): 89-104.

Brenchley, P.J., Marshall, C. R., Carden, G. A. F., Robertson, D. B. R., Long, D. G. F., Meidla, T., Hints, L., Anderson, T. F., 1994. Bathymetric and isotopic evidence for a short-lived Late Ordovician glaciation in a greenhouse period. Geology, 22: 295-298.

Brenchley, P.J. and Newall, G., 1980. A facies analysis of Upper Ordovician regressive sequences in the Oslo region, Norway; a record of glacio-eustatic changes. Palaeogeography, Palaeoclimatology, Palaeoecology, 31(1): 1-38.

Brimberry, D.L., 1984. Depositional Environments, Conodont Biostratigraphy, and Porosity Diagenesis of Montoya Group (Upper Ordovician), Sacramento Mountains, South-Central Mexico: Abstract. AAPG Bulletin, 68(4): 457-458.

Brown, D.A., Campbell, K.S.W. and Crook, K.A.W., 1968. The geological evolution of Australia & New Zealand. Oxford, New York, Pergamon Press, 409 pp.

Bruchert, V., Delano, J.W. and Kidd, W.S.F., 1994. Fe- and Mn-enrichment in Middle Ordovician hematitic argillites preceding black shale and flysch deposition: the Shoal Arm Formation, north-central Newfoundland. Journal of Geology, 102(march): 197-214.

Burchfiel, B.C., 1964. Precambrian and Paleozoic Stratigraphy of Specter Range Quadrangle, Nye County, Nevada. AAPG Bulletin, 48(1): 40-56.

Buschbach, T.C., 1964. Cambrian and Ordovician strata of northeastern Illinois. Illinois State Geological Survey Report of Investigations, 218: 90.

Byers, C.W. and Stasko, L.E., 1978. Trace Fossils and Sedimentologic Interpretation-- McGregor Member of Platteville Formation (Ordovician) of Wisconsin. Journal of Sedimentary Petrology, 48(4): 1303-1309.

Calvert, S.E., 1966. Accumulation of diatomaceous silica in sediments of the . Bull. geol. Soc. Am., 77(569-596).

Calvert, S.E., 1974. Deposition and diagenesis of silica in marine sediments. In: K.J. Hsü and H.C. Jenkyns (Editors), Pelagic sediments: on land and under sea. International Association of sedimentologists; Blackwell Science, Inc., pp. 301 - 326. 192

Calvert, S.E., 1983. Sedimentary geochemistry of silicon. In: S.R. Aston (Editor), Silicon geochemistry and biogeochemistry. Academic Press, Inc., New York, pp. 143- 186.

Calvert, S.E., 1987. Oceanographic controls on the accumulation of organic matter in marine sediments. In: J. Brooks and A.J. Fleet (Editors), Marine Petroleum Source Rocks. Geological Society Special Publications. Blackwell Scientific Publications, Oxford, pp. 137-151.

Campbell, J.A. and Suneson, N.H., 1989. Petroleum Geology of Ouachita Uplift Region of Oklahoma: Abstract. AAPG Bulletin, 73(8): 1044.

Carrera, M.G. and Rigby, J.K., 1999. Biogeography of Ordovician sponges. Journal of Paleontology, 73(1): 26-37.

Carroll, D., 1959. Zircon from a bentonite bed in Martinsburg shale (Ordovician) at Fisher's Hill, Virginia. GSA Bulletin, 70: 223-224.

Carson, G.A., 1991. Silicification of fossils. In: P.A. Allison and D.E.G. Briggs (Editors), Taphonomy: releasing the data locked in the fossil record. Topics in Geobiology, New York, pp. 455-499.

Casey, R.E., 1993. Radiolaria, Fossil prokaryotes and protists. Boston : Blackwell Scientific Publications, pp. 249-284.

Chang, F., Zhuang, Lihua, Li, Tiegang, Yan, Jun, Cao, Qiyuan, Cang, Shuxi,l 2003. Radiolarian fauna in surface sediments of the northeastern East China Sea. Marine Micropaleontology, 48: 169-204.

Churkin, M.J., 1962. Facies Across Paleozoic Miogeosynclinal Margin of Central Idaho. AAPG Bulletin, 46(5): 569-591.

Churkin, M.J., 1963. Graptolite Beds in Thrust Plates of Central Idaho and Their Correlation with Sequence in Nevada. AAPG Bulletin, 47(8): 1611-1623.

Churkin, M.J., 1975. Basement Rocks of Barrow Arch, Alaska, and Circum- Paleozoic Mobile Belt. AAPG Bulletin, 59(3): 451-456.

Clayton, C.J., 1986. The chemical environment of flint formation in Upper Cretaceous chalks. In: G.d.G. Sieveking and M.B. Hart (Editors), The scientific study of flint and chert. Cambridge Univ. Press, pp. 43-54.

193

Cohee, G.V., 1948. Cambrian and Ordovician Rocks in Michigan Basin and Adjoining Areas. AAPG Bulletin, 32(8).

Cole, T., 1941. Subsurface Study of Ellenburger Formation in West Texas: ABSTRACT. AAPG Bulletin, 25(5): 934-935.

Condie, K.C. and Snansieng, S., 1971. Petrology and geochemistry of the Duzel (Ordovician) and Gazelle (Silurian) Formations, northern California. Journal of Sedimentary Petrology, 41(3): 741-751.

Cone, M., Mitchell, C.E., Goldman, D., Maletz, J. and Janousek, H., 2002. A revised Ordovician graptolite biostratigraphy of the Phi Kappa Formation at trail creek, central Idaho. 2002 Denver Annual Meeting (October 27-30, 2002).

Coniglio, M., 1987. Biogenic chert in the Cow Head Group (Cambro-ordovician), western Newfoundland. Sedimentology, 34: 813-823.

Conley, C.D., 1977. Origin of Distorted Ooliths and Pisoliths. Journal of Sedimentary Petrology, 47(2): 554-564.

Conway, S.M., 1994. Early metazoan evolution: first steps to an integration of molecular and morphological data. In: S. Bengtson (Editor), Early life on earth / Nobel Symposium no. 84. Columbia University Press, New York, pp. 450-459.

Cowen, R., 2000. The history of life. Blackwell Science, 432 pp.

Craig, W.W., 1993. Stratigraphy of the Cason Shale, the Brassfield, St. Clair, and Lafferty limestones, and the Penters Chert (Upper Ordovician-Lower Devonian) in the Arkansas Ozarks. In: K.S. Johnson, editor. (Editor), Special Publication - Oklahoma Geological Survey. 93-4. Oklahoma Geological Survey, Oklahoma, pp. 135-147.

Crimes, T.P., 1974. Colonisation of the early ocean floor. Nature (London), 248(5446): 328-330.

Crimes, T.P., 1992. The record of trace fossils across the Proterozoic-Cambrian boundary. In: J.H. Lipps and P.W. Signor (Editors), Origin and early evolution of the Metazoa. Topics in geobiology. Plenum Press, New York, pp. 177-202.

Dalla Salda, L.H., Cingolani, C.A. and Varela, R., 1992. Early Paleozoic orogenic belt of the Andes in southwestern South America; result of Laurentia-Gondwana collision? Geology (Boulder), 20(7): 617-620.

194

Danelian, T. and Clarkson, E.N.K., 1998. Ordovician Radiolaria from bedded cherts of the Southern Uplands. Scottish Journal of Geology, 34(2): 133-137.

Davis, E.F., 1918. The radiolarian cherts of the Franciscan group. Univ. Calif. Publs Bull. Dep. Geol., 11: 235-432.

De La Rocha, C.L., Brzezinski, M.A. and DeNiro, M.J., 1997. Fractionation of silicon isotopes by marine diatoms during biogenic silica formation. Geochimica and Cosmochimica Acta, 61(23): 5051-5056.

De Wever, P., 1994. Radiolarians and radiolarites. Comptes Rendus de l'Academie des Sciences, Serie II. Sciences de la Terre et des Planetes, 319(5): 513-526.

De Wever, P., Azema, J. and Fourcade, E., 1994. Radiolaires et radiolarites; production primaire, diagenese et paleogeographie. Bulletin des Centres de Recherches Exploration-Production Elf-Aquitaine, 18(1): 315-379.

De Wever, P., Dumitrica, P., Caulet, J.P., Nigrini, C. and Caridroit, M., 2001. Radiolarians in the Sedimentary Record. Gordon and Breach Science Publisher, 533 pp.

DeMaster, D.J., 1981. The supply and accumulation of silica in the marine environment. Geochimica and Cosmochimica Acta, 45: 1715-1732.

Denison, R.E., 1997. Contrasting sedimentation inside and outside the southern Oklahoma Aulacogen during the Middle and Late Ordovician. In: K.S. Johnson, editor. (Editor), Simpson and Viola groups in the southern Midcontinent, 1994 symposium; proceedings. Circular - Oklahoma Geological Survey, University of Oklahoma, pp. 39-47.

Dietz, R.S. and Holden, J.C., 1966. Deep-Sea Deposits in but not on the continents. AAPG Bulletin, 50(2): 351-362.

Dolly, E.D. and Busch, D.A., 1972. Stratigraphic, Structural, and Geomorphologic Factors Controlling Oil Accumulation in Upper Cambrian Strata of Central Ohio. AAPG Bulletin, 56(12): 2335-2368.

Donovan, D.T. and Jones, E.J.W., 1981. Causes of world-wide changes in sea level. Journal of the Geological Society, 136: 187-192.

Drever, J.I., 1997. The geochemistry of natural waters - Surface and groundwater environments. Prentice Hall, Inc., New Jersey, 436 pp.

195

Drew, L.J., Berger, B.R. and Kurbanov, N.K., 1996. Geology and structural evolution of the Muruntau gold deposit, Kyzylkum desert, Uzbekistan. Ore Geology Reviews, 11(4): 175-196.

Dumitrica, P., 1970. Cryptocephalic and cryptothoracic Nassellaria in some Mesozoic deposits of Romania. Revue Roumaine de Geologie, Geophysique et Geographie, Serie de Geologie, 14(1): 45-124.

Duxbury, A.B., Duxbury, A.C. and Sverdrup, K.A., 2002. Fundamentals of oceanography. McGraw-Hill Higher Education, 344 pp.

Eastoe, C.J. and Gustin, M.M., 1996. Volvanogenic massive sulfide deposits and anoxia in the Phanerozoic oceans. Ore Geology Reviews, 10: 176-197.

Ebneth, S., Shields, G.A., Veizer, J., Miller, J.F. and Shergold, J.H., 2001. High- resolution strontium isotope stratigraphy across the Cambrian-Ordovician transition. Geochimica and Cosmochimica Acta, 65(14): 2273-2292.

Edson, F.C., 1927. Ordovician Correlations in Oklahoma. AAPG Bulletin, 11(9): 967- 975.

Epis, R.C. and Gilbert, C.M., 1957. Early Paleozoic Strata in Southeastern Arizona. AAPG Bulletin, 41(10): 2223-2242.

Erdtmann, B.-D., Kley, J. and J, M., 1995. Ordovician basin dynamics and new graptolite data from the Tarija region, Eastern Cordilliera, South Bolivia. In: J.D. Cooper, M.L. Droser and S.C. Finney (Editors), Ordovician odyssey; short papers for the Seventh international symposium on the Ordovician System. Field Trip Guidebook - Pacific Section, (SEPM), Fullerton, pp. 498.

Feray, D.E., Oetking, P. and Renfro, H.B., 1968. Geological Highway Map Pacific southwest region California-Nevada. AAPG, Tulsa, Oklahoma.

Fergusson, C.L. and Fanning, C.M., 2002. Late Ordovician stratigraphy, zircon provenance and tectonics, Lachlan Fold Belt, southeastern Australia*. Australian Journal of Earth Sciences, 49(3): 423-436.

Finks, R.M., 1960. Late Paleozoic sponge faunas of the Texas region--The siliceous sponges. Bulletin of the American Museum of Natural History, 120: 1-160, art. 1.

196

Finks, R.M., 1970. The evolution and ecologic history of sponges during paleozoic times. In: W.G. Fry (Editor), The biology of porifera. Symposia of the Zoological Society of London, no. 25. Published for the Zoological Society of London by Academic Press, pp. 512.

Finks, R.M., 1983. Fossil Hexactinellida. In: T.W. Broadhead (Editor), Sponges and spongiomorphs : notes for a short course organized by J.K. Rigby and C.W. Stearn. Knoxville, TN : University of Tennessee, Dept. of Geological Sciences, pp. 101-129.

Finney, S.C., 1997. Ordovician sea-level changes recorded in deep-water, continental- margin facies of North America. In: K.S. Johnson, editor. (Editor), Simpson and Viola groups in the southern Midcontinent, 1994 symposium; proceedings. Circular - Oklahoma Geological Survey. Circular - Oklahoma Geological Survey, pp. 103-110.

Finney, S.C., Achab, Aicha, Berry, William B. N., Cooper, John D., Jacobson, Stephen R., Noble, Paula J., Ripperdan, Robert L., Soufiane, Azzedine, Sweet, Walter C., 1999. Late Ordovician mass extinction; a new perspective from stratigraphic sections in central Nevada. Geology (Boulder), 27(3): 215-218.

Finney, S.C. and Bartolini, C., 1993. The Record of Ordovician Sea Level Changes in Outer Settings of North America Abstract. AAPG Bulletin, 77(4): 696.

Fischer, A.G., 1984. Biological innovations and the sedimentary record. In: H.D. Holland and A.F. Trendall (Editors), Patterns of Change in Earth Evolution. Physical, Chemical, and Earth Sciences Research Reports. Springer-Verlag, Berlin, pp. 145-157.

Fischer, A.G. and Arthur, M.A., 1977. Secular variations in the pelagic realm. In: H.E. Cook and P. Enos (Editors), Deep-water carbonate environments. Society of Economic Paleontologists and Mineralogists Special Publications. Society of Economic Paleontologists and Mineralogists, Tulsa, Oklahoma, pp. 19-50.

Fisher, I.S., 1968. Interrelation of Mineralogy and Texture Within an Ordovician Section in Central Kentucky. Journal of Sedimentary Petrology, 38(3): 775-784.

Fisher, J.H. and Barratt, M.W., 1985. Exploration in Ordovician of Central Michigan Basin. AAPG Bulletin, 69(12): 2065-2076.

197

Floyd, J., 1996. Lithostratigraphy of the Ordovician rocks in the Southern Uplands: Crawford Group, Moffat Shale Group, Leadhills Supergroup. Transactions of the Royal Society of Edinburgh-Earth Sciences, 86(3): 153-165.

Folk, R.L. and Pittman, J.S., 1971. Length-slow Chalcedony: A New Testament for Vanished Evaporites. Journal of Sedimentary Petrology, 41(4): 1045-1058.

Ford, D. and Golonka, J., 2003. Phanerozoic paleogeography, paleoenvironment and lithofacies maps of the circum-Atlantic margins. Marine and petroleum geology, 20((3-4)): 249-285.

Fortey, R.A., 1984. Global earlier Ordovician transgressions and regressions and their biological implications. In: D.L. Bruton (Editor), Paleontological Contributions from the University of Oslo. Paleontologisk Museum, Oslo, pp. 37-50.

Frakes, L.A., Francis, J.E. and I., S.J., 1992. Climate modes of the Phanerozoic. The history of the Earth's climate over the past 600 million years. Cambridge University Press, Cambridge, 274+XI pp.

François, L.M., Walker, J.C.G. and Opdyke, B.N., 1993. The histrory of Global weathering and the chemical evolution of the Ocean-atmosphere system. In: E. Takahashi, R. Jeanloz and D. Rubie (Editors), Evolution of the Earth and Planets, Geophysical Monograph. American Geophysical Union : International Union of Geodesy and Geophysics, pp. 143-159.

Frøhlich, H. and Barthel, D., 1997. Silica uptake of the marine sponge Halichondria panicea in Kiel Bight. Marine Biology, 128(1): 115-125.

Gaffin, S., 1987. Ridge volume dependence on seafloor generation rate and inversion using long term sealevel change. American Journal of Science, 287(6): 596-611.

Gallardo, J. and Blackwell3, D.D., 1999. Thermal Structure of the Anadarko Basin. AAPG Bulletin, 83(2): 333-361.

Gao, G. and Land, L.S., 1991. Early Ordovician Cool Creek Dolomite, Middle Arbuckle Group, Slick Hills, Sw Oklahoma, U.S.A.: Origin and Modification. Journal of Sedimentary Petrology, 61(2): 161-173.

Garrison, R.E., 1974. Radiolarian cherts, pelagic limestones, and igneous rocks in eugeosynclinal assemblages. In: I.K.J. Hsü and H.C. Jenkyns (Editors), Pelagic sediments: on land and under sea, International Association of sedimentologists, Special Publication 1. Blackwell Science, Inc., pp. 367 - 399.

198

Gehling, J.G. and Rigby, J.K., 1996. Long expected sponges from the Neoproterozoic Ediacara fauna of south Australia. Journal of Paleontology, 70(2): 185-195.

Gentzis, T., Freitas, T.d., Goodarzi, F., Melchin, M. and Lenz, A., 1996. Thermal Maturity of Lower Paleozoic Sedimentary Successions in Arctic Canada. AAPG Bulletin, 80(7): 1065-1083.

Gerhard, L.C., 1967. Paleozoic Geologic Development of Canon City Embayment, Colorado. AAPG Bulletin, 51(11): 2260-2280.

Gerhard, L.C., 1974. Redescription and New Nomenclature of Manitou Formation, Colorado: Geological Notes. AAPG Bulletin, 58(7): 1397-1402.

Gibbs, M.T., Barron, E.J. and Kump, L.R., 1997. An atmospheric pCO2 threshold for glaciation in the Late Ordovician. Geology, 25(5): 447-450.

Girty, G.H., Al-Riyami, R.K., Johnson, D., Knaack, C. and Ridge, D.L., 1996. Provenance and depositional setting of Paleozoic chert and argillite, Sierra Nevada, California. Journal of Sedimentary Research, 66(1): 107-118.

Grammer, G.M., 1983. Diagenesis of Viola Limestone (Middle and Upper Ordovician), Southeastern Arbuckle Mountains, Oklahoma: Abstract. AAPG Bulletin, 67(3): 473-474.

Gregg, J.M. and Sibley, D.F., 1984. Epigenetic Dolomitization and the Origin of Xenotopic Dolomite Texture. Journal of Sedimentary Petrology, 54(3): 908-931.

Gregory, R.T., 1991. Oxygen isotope history of seawater revisited: timescales for boundary event changes in the oxygen isotope composition of seawater. In: T. H.P., O.N. J.R. and K. I.R. (Editors), Stable isotope geochemistry: A Tribute to Samuel Epstein. Geochemical Society Special Publication, pp. 65–76.

Grenne, T. and Slack, J.F., 2003. Paleozoic and Mesozoic silica-rich seawater; evidence from hematitic chert (jasper) deposits. Geology, 31(4): 319-322.

Grotzinger, J.P., Bowring, S.A., Kaufman, A.J. and Saylor, B.Z., 1995. Biostratigraphic and geochronologic constraints on early animal evolution. Science, 270(5236): 598-604.

Grunau, H.R., 1965. Radiolarian cherts and associated rocks in space and time. Ecologae Geologicae Helvetiae, 58: 145-208.

199

Harland, W.B., Armstrong, Richard L., Cox, Allen V., Craig, Lorraine E., Smith, Alan G., Smith, David G.1990. A geologic time scale, 1989. Cambridge Univ. Press, 263 pp.

Harlton, B.H., 1963. Frontal Wichita Fault System of Southwestern Oklahoma. AAPG Bulletin, 47(8): 1552-1580.

Harlton, B.H., 1966. Relation of Buried Tishomingo Uplift to Ardmore Basin and , Southeastern Oklahoma. AAPG Bulletin, 50(7): 1365-1374.

Harris, A.G., Dumoulin, J.A., Repetski, J.E. and Carter, C., 1995. Correlation of Ordovician rocks of Northern Alaska. In: J.D. Cooper, M.L. Droser and S.C. Finney (Editors), Ordovician odyssey : short papers for the Seventh International Symposium on the Ordovician System. The Pacific Section Society for Sedimentary Geplogy, pp. 21-26.

Harris, L.D., 1958. Syngenetic chert in the Middle Ordovician Hardy Creek limestone of southwest Virginia. Journal Of Sedimentary Petrology, 28(2): 205-208.

Hartman, W.D., 1983. Modern Hexactinellida. In: T.W. Broadhead (Editor), Sponges and spongiomorphs : notes for a short course organized by J.K. Rigby and C.W. Stearn. Knoxville, TN : University of Tennessee, Dept. of Geological Sciences, pp. 90-95.

Hays, J.D. and Pitman, W.C., III, 1973. Lithospheric Plate Motion, Sea Level Changes and Climatic and Ecological Consequences. Nature (London), 246(5427): 18-22.

Heath, G.R., 1973. Cherts from the Eastern Pacific. In: T.H.v. Andel and G.R. Heath (Editors), Initial Reports of the Deep Sea Drilling Project. U.S. Government Printing Office, Washington, pp. 609-613.

Heath, G.R., 1974. Dissolved silica and deep-sea sediments. In: W.W. Hay (Editor), Studies in Paleo-oceanography. Society of Economic Paleontologists and Mineralogists Special Publications, 20. Society of Economic Paleontologists and Mineralogists, Tulsa, Oklahoma, pp. 77-93.

Heath, G.R. and Moberly, R., 1971. Cherts from the Western Pacific, Leg 7, Deep Sea Drilling Project. In: E.L.W.e. al. (Editor), Initial Reports of the Deep Sea Drilling Project. U.S. Government Printing Office, Washington, pp. 991-1007.

Hein, J.R. and Parrish, J.T., 1987. Distribution of siliceous deposits in space and time. In: J.R. Hein and W. Walker (Editors), Siliceous sedimentary rock-hosted ores and petroleum. Van Nostrand Reinhold Co., New York, pp. 10-57. 200

Herrmann, A., D., Patzkowski, M., E. and Pollard, D., 2003. Obliquity forcing with 8-12 times preindustrial levels of atmospheric pCO2 during the Late Ordovician glaciation. Geology, 31(6): 485-488.

Herrmann, A.D., Patzkowsky, M.E. and Pollard, D., 2004. The impact of paleogeography, pCO (sub 2) , poleward ocean heat transport and sea level change on global cooling during the Late Ordovician. Palaeogeography, Palaeoclimatology, Palaeoecology, 206(1-2): 59-74.

Hobson, J.P., 1957. Lower Ordovician (Beekmantown) Succession in Berks County, Pennsylvania. AAPG Bulletin, 41(12): 2710-2722.

Holland, S.M. and Patzkowsky, M.E., 1997. Distal Orogenic Effects on Peripheral Bulge Sedimentation: Middle and Upper Ordovician of the Nashville Dome. Journal of Sedimentary Research, Section B: Stratigraphy and Global Studies, 67(2): 250- 263.

Horodyski, R.J., Bauld, J., Lipps, J.H. and Mendelson, C.V., 1992. Preservation of prokaryotes and organic-walled and calcareous and siliceous protists. In: J.W. Schopf and C. Klein (Editors), Proterozoic biosphere; a multidisciplinary study. Cambridge Univ. Press, pp. 185-193.

Hubert, J.F., 1966. Sedimentary history of Upper Ordovician geosynclinal rocks, Girvan, Scotland. Journal of Sedimentary Petrology, 36(3): 677-699.

Huff, W.D., Bergstrom, S.M. and Kolata, D.R., 1992. Gigantic Ordovician volcanic ash fall in North America and Europe; biological, tectonomagmatic, and event- stratigraphic significance. Geology (Boulder), 20(10): 875-878.

Huff, W.D., Bergstrom, S.M., Kolata, D.R. and Zhang, Y.S., 1996. Large-magnitude Middle Ordovician volcanic ash falls in North America and Europe; dimensions, emplacement and post-emplacement characteristics. Journal of Volcanology and Geothermal Research, 73(3-4): 285-301.

Huff, W.D. and Kolata, D.R., 1990. Correlation of the Ordovician Deicke and Millbrig K-Bentonites Between the Mississippi Valley and the Southern Appalachians. AAPG Bulletin, 74(11): 1736-1747.

Iams, W. and Stevens, R.K., 1988. Radiolaria and other siliceous of the Cow Head Group (Upper Cambrian - Middle Ordovician) of western Newfoundland. Geologica and Palaeontologica, 22: 192-193.

201

Igo, H., Adachi, S., Furutani, H. and Nishiyama, H., 1980. Ordovician fossils first discovered in Japan. Proceedings of the Japan Academy. Series B: Physical and Biological Sciences, 56(8): 499-503.

Iwata, K., Schmidt, B., Leitch, E., Allan, A. and Watanabe, T., 1995. Ordovician microfossils from the Ballast formation (Girilambone group) of new-south-Wales. Australian Journal of Earth Sciences, 42(4): 371-376.

Jablonski, D., Sepkoski, J.J.J., Bottjer, D.J. and Sheenan, P.M., 1983. Onshore-offshore patterns in the evolution of Phanerozoic shelf communities. Science, 222: 1123- 1125.

Jackson, D.E. and Lenz, A.C., 1962. Zonation of Ordovician and Silurian graptolites of northern Yukon, Canada. AAPG Bulletin, 46: 30-45.

Jones, B. and Desrochers, A., 1992. Shallow platform carbonates. In: R.G. Walker and N.P. James (Editors), facies Models. Response to sea level change. Geological Association of Canada, pp. 454.

Jones, D.L. and Murchey, B.L., 1986. Geologic significance of Paleozoic and Mesozoic radiolarian chert. Annual Review of Earth and Planetary Sciences, 14: 455-492.

Kamatani, A., 1982. Dissolution rates of silica from diatoms decomposing at various temperatures. Marine Biology, 68: 91-96.

Kastner, M., 1981. Authigenic silicates in deep-sea sediments: Formation and diagenesis. In: C. Emiliani (Editor), Oceanic lithosphere. The Sea. Research Program: DSDP @Deep Sea Drilling Project. NY : John Wiley and Sons, New York, pp. 915-980.

Kastner, M., Keene, J.B. and Gieskes, J.M., 1977. Diagenesis of siliceous oozes - I. Chemical controls on the rate of opal-A to opal-CT transformation - an experimental study. Geochimica and Cosmochimica Acta, 41: 1041-1059.

Kay, M., 1948. Summary of Middle Ordovician Bordering Allegheny Synclinorium. AAPG Bulletin, 32(8): 1397-1416.

Kay, M., 1967. Stratigraphy and Structure of Northeastern Newfoundland Bearing on Drift in North Atlantic. AAPG Bulletin, 51(4): 579-600.

Kennett, J., 1982. Marine geology. Prentice-hall, Inc., Englewood Cliffs, N.J, 813 pp.

202

Kidder, D.L., 2000. Systemic responses to biogenic drawdown of dissolved ocean silica in the early Paleozoic. Abstracts with Programs - Geological Society of America, 32(7): 66.

Kidder, D.L. and Erwin, D.H., 2001. Secular distribution of biogenetic silica through the Phanerozoic: comparison of silica-replaced fossils and bedded cherts at the series level. The Journal of Geology, 109: 509-522.

Kidder, D.L. and Mumma, S.A., 2003. Silica-replaced iilotes, bedded shelf cherts and Paleozoic changes in the silica cycle. Sedimentary Geology, 162: 159-166.

Kim, D.H. and Choi, D.K., 2000. Lithostratigraphy and biostratigraphy of the Mungok Formation (Lower Ordovician), Yongwol, Korea. Geosciences Journal (Seoul), 4(4): 301-311.

King, P.B., 1926. The Geologic Structure of a Portion of the Glass Mountains of West Texas. AAPG Bulletin, 10(9): 877.

Knauth, L.P., 1979. A model for the origin of chert in limestone. Geology, 7: 274-277.

Knoll, A.H. and Lipps, J.H., 1993. Evolutionary history of prokaryotes and protists. In: J.H. Lipps (Editor), Fossil Prokaryotes and Protists. Boston : Blackwell Scientific Publications, pp. 19-30.

Kopaska-Merkel, D.C. and Friedman, G.M., 1989. Petrofacies analysis of carbonate rocks: example from lower Paleozoic Hunton Group of Oklahoma and Texas. AAPG Bulletin, 73(11): 1289-1306.

Kozur, H.W., Mostler, H. and Repetski, J.E., 1996. "Modern" siliceous sponges from the lowermost Ordovician (early Ibexian – early Tremadocian) Windfall Formation of the Antelope Range, Eureka County, Nevada, U.S.A. Geologisch- Paläontologische Mitteilungen Innsbruck, 21: 201-221.

Kruglikova, S.B., 1993. Observations on the distribution of polycystine Radiolaria marine sediments (mainly at high taxonomic levels). In: J.R. Blueford and B. Murchey (Editors), Radiolaria of giant and subgiant fields in Asia: Nazarov memorial volume. Micropaleontology Press, New York, pp. 17-21.

Kruse, P.D., 1983. Middle Cambrian "Archaeocyathus" from the Georgina Basin is an anthaspidellid sponge. Alcheringa, 7(1): 49-58.

203

Kump, L.R, Arthur, M. A., Patzkowski, M. E., Gibbs, M. T., Pinkus, D. S., Sheenan, P. M., 1999. A weathering hypothesis for glaciation at high atmospheric pCO2 during the Late Ordovician. Palaeogeography, Palaeoclimatology, Palaeoecology, 152: 173-187.

Kusky, T.M. and Kidd, W.S.F., 1996. Tectonic implications of early Silurian thrust imbrication of the northern exploits subzone, central Newfoundland. Journal of Geodynamics, 22(3-4): 229-265.

La Porta, P.C., 1998a. Chertification processes and silica sources; examples from the Central Appalachians. Geological Society of America Abstracts with Programs, 30(7): 333.

La Porta, P.C., 1998b. Diagenesis of Cambro-Ordovician cherts. Geological Society of America Abstracts with Programs, 30(1): 31.

Lancelot, Y., 1973a. Formation on deep-sea chert: role of he sedimentary environment (Abstract). In: K.J. Hsü and H.C. Jenkyns (Editors), Pelagic sediments: on land and under sea. International Association of sedimentologists, Blackwell Science, Inc., pp. 447.

Lancelot, Y., 1973b. Chert and silica diagenesis in sediments from the central Pacific. Init. Repts. DSDP, 17: 377-406.

Langenheim, R.L., Barnes, J.A., Delise, K.C., Ross, W.A. and Stanton, J.M., 1956. Middle and Upper(?) Ordovician Rocks of Independence Quadrangle, California. AAPG Bulletin, 40(9): 2081-2097.

Langenheim, R.L., Carss, B.W., Kennerly, J.B., McCutcheon, V.A. and Waines, R.H., 1962. Paleozoic Section in Arrow Canyon Range, Clark County, Nevada. AAPG Bulletin, 46(5): 592-609.

Larson, E.R., 1951. Stratigraphy of Plattin Group, Southeastern Missouri. AAPG Bulletin, 35(9): 2041-2075.

Latham, J.W., 1968. Petroleum Geology of Arbuckle Group (Ordovician), Healdton Field, Carter County, Oklahoma. AAPG Bulletin, 52(1): 3-20.

Lecuyer, C. and Allemand, P., 1999. Modelling of the oxygen isotope evolution of seawater; implications for the climate interpretation of the delta (super 18) O of marine sediments. Geochimica et Cosmochimica Acta, 63(3-4): 351-361.

204

Lee, Y.I. and Friedman, G.M., 1987. Deep-Burial Dolomitization in the Ordovician Ellenburger Group Carbonates, West Texas and Southeastern New Mexico. Journal of Sedimentary Petrology, 57(3): 544-557.

Leggett, J.K., Cocks, L.R.M., Mckerrow, W.S. and Rickards, R.B., 1981. Periodicity in the early Palaeozoic marine realm. Journal of the Geological Society of London, 138(Part 2): 167-176.

Lewin, J.C., 1961. The dissolution of silica from diatom walls. Geochimica and Cosmochimica Acta, 21: 182-198.

Li, C.-W., Junyuan, C. and Hua, T.-E., 1998. Precambrian sponges with cellular structures. Science, 279(5352): 879-882.

Li, J., 1995. A study of the effects of organic matter on silica diagenesis using a geochemical modeling code. Master of Science Thesis, University of Montana, Missoula, MT, 58 pp.

Lindström, M., 1984. Late Ordovician climate based on the study of carbonate rocks. In: D. Bruton, L. (Editor), Aspects of the Ordovician system. Palaeontological contributions from the University of Oslo. Universitetsforlaget, Oslo, pp. 81-88.

Lipps, J.H., 1970. Plankton evolution. Evolution, 24(1): 1-22.

Lipps, J.H., 1992. Stratigraphic distribution of skeletonized Protists. In: J.W. Schopf and C. Klein (Editors), The Proterozoic biosphere : a multidisciplinary study. Cambridge ; New York : Cambridge University Press, pp. 499.

Liu, B., Rigby, J., Jiang, Y. and Zhu, Z., 1997. Lower Ordovician lithistid sponges from the eastern Yangtze Gorge area, Hubei, China. Journal of Paleontology, 71(2): 194-207.

Loi, A. and Dabard, M., 2002. Controls of sea level fluctuations on the formation of Ordovician siliceous nodules in terrigenous offshore environments. Sedimentary Geology, 153(3-4): 65-84.

Lyons, P. and Percival, I.G., 2002. Middle to Late Ordovician age for the Jindalee Group of the Lachlan Fold Belt, New South Wales: conodont evidence and some tectonic implications. Australian Journal of Earth Sciences, 49(5): 801-808.

Mackenzie, F.T. and Pigott, J.D., 1981. Tectonic controls of Phanerozoic sedimentary rock cycling. Journal of the Geological Society of London, 138(part 2): 183-196.

205

Maldonado, M., Carmona, M.C., Uriz, M.J. and Cruzado, A., 1999. Decline in Mesozoic reef-building sponges explained by silicon limitation. Nature, 401(october): 785 - 788.

Maliva, R.G., Knoll, A.H. and Siever, R., 1989. Secular change in chert distribution: a reflection of evolving biological participation in the silica cycle. Palaios, 4: 519- 532.

Maliva, R.G. and Siever, R., 1989. Nodular chert formation in carbonate rocks. Journal of Geology, 97: 421-433.

Matheney, R.K. and Knauth, L.P., 1993. New isotopic temperature estimates for early silica diagenesis in bedded cherts. Geology, 21(6): 519-522.

Matsuoka, A. and Anderson, O.R., 1992. Experimental and observational studies of rariolarian physiological ecology: 5. Temperature and salinity tolerance of Dictyocoryne truncatum. Marine Micropaleontology, 19: 299-313.

Mazzullo, S.J. and Friedman, G.M., 1975. Conceptual Model of Tidally Influenced Deposition on Margins of Epeiric Seas: Lower Ordovician (Canadian) of Eastern New York and Southwestern Vermont. AAPG Bulletin, 59(11): 2123-2141.

McBride, E.F., 1962. Flysch and associated beds of the Martinsburg Formation (Ordovician), central Appalachians. Journal of Sedimentary Petrology, 32(1): 39- 91.

McBride, E.F., 1969. Stratigraphy and sedimentology of the Fort Pena Formation. In: T. Dallas, Dallas Geol. Soc. (Editor), guidebook to the stratigraphy, sedimentary structures and origin of flysch and pre-flysch rocks of the Marathon basin, Texas- -AAPG-SEPM Ann. Mtg. 1969, dallas, pp. 43-46.

McCracken, E., 1955. Correlation of Insoluble Residue Zones of Upper Arbuckle of Missouri and Southern Kansas. AAPG Bulletin, 39(1): 49-59.

McFarlan, A.C. and White, W.H., 1948. Trenton and Pre-Trenton of Kentucky. AAPG Bulletin, 32(8): 1627-1646.

Mehl, D. and Lehnert, O., 1997. Cambro-Ordovician sponge spicules assemblages in the Ordovician of the Argentine Precordillera and paleoenvironmental. Neues Jahrbuch für Geologie Paläontologie, 204: 204-246.

Mellen, F.F., 1974. Possible Ordovician Carbonate Reservoirs in Mississippi: Geologic Notes. AAPG Bulletin, 58(5): 870-876. 206

Miller, A.I., 1997. Dissecting global diversity patterns: Examples from the Ordovician Radiation. Annual Review of Ecology and Systematics, 28: 85-104.

Miller, A.I. and Mao, S., 1995. Association of orogenic activity with the Ordovician radiation of marine life. Geology (Boulder), 23(4): 305-308.

Miller, M.F., 1977. Using Biogenic Structures to Interpret Sharp Lithologic Boundaries: An Example from the Lower Devonian of New York. Journal of Sedimentary Petrology, 52(3): 887-895.

Miskell, K.J., Brass, G.W. and Harrison, C.G.A., 1985. Global patterns in opal deposition from Late Cretaceous to late Miocene. AAPG Bulletin, 69(6): 996-1012.

Morris, R.J., 1987. The formation of organic-rich deposits in two deep-water marine environments. In: J. Brooks and A.J. Fleet (Editors), Marine Petroleum Source Rocks. Geological Society Special Publications. Blackwell Scientific Publications, Oxford, pp. 153-166.

Moshier, S.O., 1986. Carbonate Platform Sedimentology, Upper Cambrian Richland Formation, Lebanon Valley, Pennsylvania. Journal of Sedimentary Petrology, 56(2): 204-216.

Muehlenbachs, K., 1998. The oxygen isotopic composition of the oceans, sediments and the seafloor. Chemical Geology, 145: 263-273.

Murata, K.J., Friedman, I. and Gleason, J.D., 1977. Oxygen isotope relations between diagenetic silica minerals in Monterey Shale, Temblor Range, California. American Journal of Science, 277(3): 259-272.

Murchey, B., Jones, D.L. and Holdsworth, B.K., 1983. Distribution, age, and depositional environments of radiolarian chert in westren North America. In: A. Iijima, J.R. Hein and R. Siever (Editors), Siliceous deposits in the Pacific region. Developments in sedimentology. Amsterdam ; New York : Elsevier Scientific Pub. Co., pp. 94-109.

Murray, J.W., Barber, R.T., Roman, M.R., Bacon, M.P. and Feely, R.A., 1994. Physical and biological controls on carbon cycling in the equatorial Pacific. Science, 266: 58-65.

Murray, R.W., 1994. Chemical criteria to identify the depositional environment of chert; general principles and applications. Sedimentary Geology, 90(3-4): 213-232.

207

Nazarov, B.B. and Ormiston, A.R., 1986. Trends in the development of Paleozoic Radiolaria. Marine Micropaleontology, 11(1-3): 3-32.

Nazarov, B.B. and Ormiston, A.R., 1988. The biostratigraphic potential of Paleozoic radiolaria. Geologica and Palaeontologica, 22: 200.

Nazarov, B.B. and Ormiston, A.R., 1993. New biostratigraphically important Paleozoic Radiolaria of Eurasia and North America. In: J.R. Blueford and B. Murchey (Editors), Radiolaria of giant and subgiant fields in Asia. New York : Micropaleontology Press, American Museum of Natural History, pp. 22-60.

Nelson, D.M. and Gordon, L.I., 1982. Production and pelagic dissolution of biogenic silica in the Southern Ocean. Geochimica et Cosmochimica Acta, 46(4): 491-501.

Nelson, D.M., Treguer, P., Brzezinski, M.A., Leynaert, A. and Queguiner, B., 1995. Production and dissolution of biogenic silica in the ocean: revised global estimates, comparison with regional data and relationship to biogenic sedimentation. Global , 9(3): 359-72.

Nikitin, I.F., Apollonov, M.K., Tzaj, D.T., Koroljov, V.G., Kim, A.I., Erina, M.V., Larin, N.M., Golikov, A.N., 1986. Ordovician system in Kazakhstan and Middle Asia : correlation charts and explanatory notes. IUGS publication; International Union of Geological Sciences, 21.

Nikitin, I.F., Frid, N.M. and Zvontsov, V.S., 1991. Paleogeography and man features of volcanicity in the Ordovician of Kazahstan and North Tien Shan. In: C.R. Barnes and S.H. Williams (Editors), Advances in Ordovician Geology. Geological Survey of Canada Papers. Minister of Supply and Services Canada, Ottawa, pp. 259-270.

Noble, P. and Aitchison, J., 2000. Early Paleozoic radiolarian biozonation. Geology, 28(4): 367-370.

Noble, P.J., 1997. Late Ordovician crises: the siliceous microfossil record. Geological Society of America Abstracts with Programs, 29(6): 356.

Noble, P.J. and Finney, S.C., 1999. Recognition of fine-scale imbricate thrusts in lower Paleozoic orogenic belts; an example from the Roberts Mountains Allochthon, Nevada. Geology (Boulder), 27(6): 543-546.

Noble, P.J., Ketner, K.B. and McClellan, W., 1997. Early Silurian radiolaria from northen Nevada, U.S.A. Marine Micropaleontology, 30: 215-223.

208

Norford, B.S, Nowland, G.S., Bolton, T.E., Elias, R.J., Jin, J., Johnston, P.A., Rohr, D.M.., 1995. Biostratigraphy of the Advance formation, a new upper middle Ordovician unit, northern Rocky Mountains, British Columbia. In: J.D. Cooper, M.L. Droser and S.C. Finney (Editors), Ordovician odyssey; short papers for the Seventh international symposium on the Ordovician System. Field Trip Guidebook - Pacific Section, Society of Economic Paleontologists and Mineralogists. Fullerton, pp. 498.

Ormiston, A.R., 1993. The association of radiolarians with hydrocarbon source rocks. In: J.R. Blueford and B. Murchey (Editors), Radiolaria of giant and subgiant fields in Asia: Nazarov memorial volume. Micropaleontology Press, New York, pp. 9-16.

Overstreet, R.B., Gregg, J.M. and Oboh-Ikuenobe, F.E., 2003. Sequence stratigraphy and depositional facies of Lower Ordovician cyclic carbonate rocks, southern Missouri, U.S.A. Journal of Sedimentary Research, 73(3): 421-433.

Palmer, A.R. and Geissman, J., 1999. 1999 Geologic Times Scale. Geological Society of America.

P'An, C.-H., 1982. Petroleum in Basement Rocks. AAPG Bulletin, 66(10): 1597-1643.

Pancost, R.D., Freeman, K.H. and Patzkowsky, M.E., 1999. Organic-matter source variation and the expression of a late Middle Ordovician carbon isotope excursion. Geology (Boulder), 27(11): 1015-1018.

Patzkowsky, M.E., Arthur, M.A., Freeman, K.H., Pancost, R.D. and Slupik, L.M., 1997. Late Middle Ordovician environmental change and extinction; harbinger of the Late Ordovician or continuation of Cambrian patterns? Geology (Boulder), 25(10): 911-914.

Patzkowsky, M.E. and Holland, S.M., 1993. Biotic response to a Middle Ordovician paleoceanographic event in eastern North America. Geology (Boulder), 21(7): 619-622.

Paull, R.A. and Paull, R.K., 1981. Newly Discovered Inlier of Ordovician Phi Kappa Formation, South-Central Idaho: Abstract. AAPG Bulletin, 65(3): 567.

Pemberton, S.G., Frey, R.W. and James, N.P., 1992. facies models; environmental and allostratigraphic significance. In: R.G. Walker and N.P. James (Editors), Facies models; response to sea level change. Geological Association of Canada, pp. 47-72.

209

Perry, W.J., Harris, A.G. and Harris, L.D., 1979. Conodont-Based Reinterpretation of Bane Dome--Structural Reevaluation of Allegheny Frontal Zone: GEOLOGIC NOTES. AAPG Bulletin, 63(4): 647-654.

Pessagno, E.A., Jr. and Newport, R.L., 1972. A technique for extracting Radiolaria from radiolarian cherts. Micropaleontology, 18(2): 231-234.

Petrushevskaya, M.G., 1977. O proishozhdeniy radiolyariy (On the origin of Radiolaria - in Russian). Zoologicheskiy Zhurnal, 56(10): 1448-1458.

Pettijohn, F.J., 1975. Sedimentary rocks. Harper & Row, Publ., 628 pp.

Philips, S.S. and Friedman, G.M., 1995. Carbonate Diagenesis and Dolomitization of Cambro-Ordovician (Sauk Sequence) Platform Strata in Central New York: Abstract. AAPG Bulletin, 79(9): 1417-1418.

Pickerill, R., K. and Williams, P.F., 1989. Deep burrowing in the Early Paleozoic deep sea: examples from the Cambrian (?) - Early Ordovician Meguma Group of Nova Scotia. Canadian Journal of Earth Sciences, 26: 1061-1068.

Pisciotto, K.A., 1981. Diagenetic trends in the siliceous facies of the Monterey Shale in the Santa Maria region, California. Sedimentology, 28(4): 547-571.

Pitt, W.D., Cohoon, R.R., Lee, H.C., Robb, M.G. and Watson, J., 1961. Ouachita Mountain Core Area, Montgomery County, Arkansas. AAPG Bulletin, 45(1): 72- 94.

Pollock, S.G., 1987. Chert Formation in an Ordovician Volcanic Arc. Journal of Sedimentary Petrology, 57(1): 75-87.

Pope, M. and Steffen, J.B., 2003. Widespread, prolonged late Middle to Late Ordovician upwelling in North America: a record of glaciation? Geology, 31(1): 63-66.

Pope, M.C. and Steffen, J.B., 2001. Upwelling in the Montoya group: evidence for prolonged Late Ordovician glaciation. GSA Abstracts with Programs, 33(6): p. A214-A215.

Pope, M.C. and Steffen, J.B., 2002. 3rd-Order Subtidal Sequences in the Upper Ordovician Montoya Group, New Mexico and West Texas: A Record of Global Glaciation? Abstract. AAPG Bulletin, 86(13 (Supplement)).

Porter, J.W. and Fuller, J.G.C.M., 1959. Lower Paleozoic Rocks of Northern Williston Basin and Adjacent Areas. AAPG Bulletin, 43(1): 124-189. 210

Poussart, P.F., Weaver, A.J. and Barnes, C.R., 1999. Late Ordovician glaciation and high atmospheric CO2: addressing an apparent paradox via a coupled model approach. Acta Universitatis Carolinae - Geologica, 43(1/2).

Pratt, B.R., Cowan, C.A. and James, N.P., 1992. Peritidal carbonates. In: R.G. Walker and N.P. James (Editors), Facies models; response to sea level change. St. Johns, NF : Geological Association of Canada, pp. 303-322.

Prokopovich, N., 1953. Silicification in the Oneota Dolomite. Journal of Sedimentary Petrology, 23(3): 174-179.

Prothero, D., 2004. Evolution of the earth. New York Mcgraw-Hill, 567 pp.

Qing, H., Barnes, C.R., Buhl, D. and Veizer, J., 1998. The strontium isotopic composition of Ordovician and Silurian brachiopods and conodonts; relationships to geological events and implications for coeval seawater. Geochimica et Cosmochimica Acta., 62(10): 1721-1733.

Qing, S.S., 1995. Dolomite reservoirs: porosity evolution and reservoir characteristics. American Association of Petroleum Geologists Bulletin, 79(2): 186-204.

Racki, G., 1999. Silica-secreting biota and mass extinctions; survival patterns and processes. Palaeogeography, Palaeoclimatology, Palaeoecology, 154(1-2): 107- 132.

Racki, G. and Cordey, F., 2000. Radiolarian palaeoecology and radiolarites: is the present the key to the past? Earth-Science Reviews, 52(1-3): 83-120.

Ragland, D.A. and Donovan, R.N., 1990. Evidence for Existence of Sabkhalike Conditions in Upper Arbuckle Group, Slick Hills, Southwestern Oklahoma: Abstract. AAPG Bulletin, 74(2): 220-221.

Ragland, D.A. and Donovan, R.N., 1991. Bedded Cherts in the Early Ordovician Arbuckle Group of Southwestern Oklahoma. Abstract. AAPG Bulletin, 75(3): 658.

Ramsay, A.T.S., 1973. A history of organic siliceous sediments in oceans. In: N.F. Hughes (Editor), Organisms and continents through time. Special Papers in Palaeontology. The palaeontological Association London, London, pp. 199-234.

Raup, D.M., 1976. Species diversity in the Phanerozoic: an interpretation. Paleobiology, 2: 289-297.

211

Reincke, T. and Barthel, D., 1997. Silica uptake kinetics of Halichondria panicea in Kiel Bight. Marine Biology, 129(4): 591-593.

Renfro, H.B. and Feray, D.E., 1989. Mid-Atlantic region Geological highway map, Kentucky-Tennessee, west Virginia-Virginia, Maryland-Delaware, North Carolina, South Carolina. AAPG, Tulsa, Oklahoma.

Renz, G.W., 1990. Ordovician Radiolaria from Nevada and Newfoundland; a comparison at the family level. Marine Micropaleontology, 15(3-4): 393-402.

Richter, F.M., Rowley, D.B. and DePaolo, D.J., 1992. Sr isotope evolution of seawater: the role of tectonics. Earth and Planetary Science Letters, 109: 11-23.

Rigby, J.K., 1983. Introduction to the Porifera. In: T.W. Broadhead (Editor), Sponges and spongiomorphs : notes for a short course organized by J.K. Rigby and C.W. Stearn. Knoxville, TN : University of Tennessee, Dept. of Geological Sciences, pp. 1-12.

Rigby, J.K. and Webby, B.D., 1988. Late Ordovician sponges from the Malongulli Formation of central New South Wales, Australia. Palaeontographica Americana, 56.

Rigby, S. and Milsom, C.V., 2000. Origins, evolution, and diversification of zooplankton. Annual Reviews Ecological Systems, 31: 293-313.

Ronov, A.B., 1982. The earth's sedimentary shell (quantitative patterns of its structure, compositions, and evolution) - Part One. International Geology Review, 24(11): 1313-1364.

Ronov, A.B., Khain, V.E., Balukhovsky, A.N. and Seslavinsky, K.B., 1980. Quantitative analysis of Phanerozoic sedimentation. Sedimentary Geology, 25: 311-325.

Ross, C.A. and Ross, J.R.P., 1995. North American Ordovician depositional sequences and correlations. In: J.D. Cooper, M.L. Droser and S.C. Finney (Editors), Ordovician odyssey : short papers for the Seventh International Symposium on the Ordovician System. The Pacific Section Society for Sedimentary Geplogy, pp. 309-314.

Ross, J.R.P. and Ross, C.A., 1992. Ordovician sea level fluctuations. In: B.D. Webby and J.R. Laurie (Editors), Global perspectives on Ordovician geology. Proceedings of the International Symposium on the Ordovician System. A.A. Balkema, pp. 327- 335.

212

Ross, M.I., 1997. PaleoGeographic Information System/Mac TM. Earth in Motion Technologies.

Ross, R.J., Jr., 1964. Relations of Middle Ordovician time and rock units in Basin Ranges, western United States. AAPG Bulletin, 48(9): 1526-1554.

Rozanov, A.Y. and Zhuravlev, A.Y., 1992. The Lower Cmbrian fossil record of the Soviet Union. In: J.H. Lipps and P.W. Signor (Editors), Origin and earlyevolution of the Metazoa. Topics in geobiology. Plenum Press, New York, pp. 205-282.

Ruedemann, P. and Redmon, H.E.R., 1927. The Turkey Mountain Lime Pools, Oklahoma. AAPG Bulletin, 11(9): 933-944.

Ryherd, T.J. and Paris, C.E., 1987. Ordovician through Silurian carbonate base-of-slope apron sequence, northern Seward Peninsula, Alaska. In: P. Weimer (Editor), Field Trip Guidebook - Pacific Section, Society of Economic Paleontologists and Mineralogists. Special Issue Alaskan North Slope geology. Los Angeles, pp. 347- 348.

Ryu, I.-C., 2002. Tectonic and stratigraphic significance of the Middle Ordovician carbonate breccias in the Ogcheon Belt, South Korea. The Island Arc, 11(3): 149- 169.

Schallreuter, R., 1989. Weitere mittelordovizische Hornsteintypen und Ostrakoden von Sylt - Further Middle Ordovician types of chert and ostracodes from Sylt, N. Germany. Neues Jahrbuch fuer Geologie und Palaeontologie. Monatshefte, 1989(4): 243-256.

Scholle, P.A., Arthur, M.A. and Ekdale, A.A., 1983. Pelagic environment. In: P.A. Scholle, D.G. Bebout and C.H. Moore (Editors), Carbonate depositional environments. AAPG Memoir, Tulsa, OK, pp. 619-691.

Sepkoski, J.J., Jr. and Sheehan, P.M., 1983. Diversification, faunal change, and community replacement during the Ordovician radiations. In: M.J.S. Tevesz and P.L. McCall (Editors), Biotic interactions in Recent and fossil benthic communities. Plenum Press, New York, pp. 673-717.

Sepkoski, J.J.J., 1995. The Ordovician radiations; diversification and extinction shown by global genus-level taxonomic data. In: J.D. Cooper, M.L. Droser and S.C. Finney (Editors), Ordovician odyssey : short papers for the Seventh International Symposium on the Ordovician System. The Pacific Section Society for Sedimentary Geplogy, Los Angeles, CA, pp. 393-396.

213

Sepkoski, J.J.J. and Miller, A.I., 1985. Evolutionary faunas and the distribution of Paleozoic marine communities in space and time. In: J.W. Valentine (Editor), Phanerozoic diversity patterns: profiles in macroevolution. Princeton University Press, Princeton, N.J., pp. 153-190.

Severson, J.L., 1950. Devonian Stratigraphy, Sunwapta Pass Area, Alberta, Canada. AAPG Bulletin, 34(9): 1826-1849.

Shepard, F.P., 1973. Submarine geology. Harper and Row, New York, 551 pp.

Shields, G.A, Carden, Giles A. F., Li, Rong-Yu, Meidla, Tonu, Rong Jiayu, Veizer, Jan., 2003. Sr, C, and O isotope geochemistry of Ordovician brachiopods; a major isotopic event around the Middle-Late Ordovician transition. Geochimica et Cosmochimica Acta, 67(11): 2005-2025.

Siever, R., 1957. The silica budget in the sedimentary cycle. The American Mineralogist, 42: 821-841.

Siever, R., 1962. Silica solubility, 0o-200oC., and the diagenesis of siliceous sediments. The Journal of Geology, 70(2): 127-149.

Siever, R., 1991. Silica in the oceans: biological-geochemical interplay. In: S.H. Schneider and P.J. Boston (Editors), Scientists on Gaia. The MIT Press, Cambridge, Massachussetts, pp. 287-295.

Siever, R., 1992. The silica cycle in the Precambrian. Geochimica and Cosmochimica Acta, 56: 3265-3272.

Sloan, D., 1976. Middle Ordovician environments of deposition, mazourka canyon, Inyo county, California. In: J.N. Moore and A.E. Fritsche (Editors), Depositional environments of Lower Paleozoic rocks in the White-Inyo Mountains, Inyo County, California. Los Angeles : Pacific Section, Society of Economic Paleontologists and Mineralogists, pp. 61-69.

Sloan, R.E., 1988a. The Deicke ash bed and the Blackriver/Trenton boundary. Abstracts with Programs - Geological Society of America, 20(5): 389.

Sloan, R.E., 1988b. A precise Ordovician time scale and the absence of cyclic extinction. Abstracts with Programs - Geological Society of America, 20(7): 105-106.

Sloss, L.L., 1963. Sequences in the cratonic interior of North America. Geological Society of America Bulletin, 74(2): 93-113.

214

Smith, G.L., Byers, C.W. and Robert H. Dott, J., 1993. Sequence Stratigraphy of the Lower Ordovician Prairie Du Chien Group on the Wisconsin Arch and in the Michigan Basin. AAPG Bulletin, 77(1): 49-67.

Sodero, D.E. and John P. Hobson, J., 1979. Depositional Facies of Lower Paleozoic Allen Bay Carbonate Rocks and Contiguous Shelf and Basin Strata, Cornwallis and Griffith Islands, Northwest Territories, Canada. AAPG Bulletin, 63(7): 1059- 1091.

Spencer, C.P., 1983. Marine biochemistry of silicon. In: S.R. Aston (Editor), Silicon geochemistry and biogeochemistry. Academic Press, Inc., New York, pp. 101- 141.

Steinberg, M., 1981. Biosiliceous sedimentation, periods and silica budget fluctuations. Oceanologica Acta, Simposium Proceedings: 149-154.

Stieglitz, R.D., 1975. Sparry White Dolomite and Porosity in Trenton Limestone (Middle Ordovician) of Northwestern Ohio: Geologic Notes. AAPG Bulletin, 59(3): 530- 533.

Stillman, C.J., 1984. Ordovician volcanicity. In: D. Bruton, L. (Editor), Aspects of the Ordovician system. Palaeontological contributions from the University of Oslo. Universitetsforlaget, Oslo, pp. 183-194.

Stokstad, E.R., 1995. Geology of the Mid-Ordovician Barrel Spring Formation, Mazourka Canyon, Inyo Mts., CA. Thesis (Master's) Thesis, University of California, Riverside, 131 pp.

Stone, C.G. and Sterling, P.J., 1962. New Lithologic Marker Horizons in Ordovician Rocks, Eastern Ouachitas of Arkansas: Geologic Notes. AAPG Bulletin, 46(3): 387-390.

Suchecki, R.K. and Hubert, J.F., 1984. Stable Isotopic and Elemental Relationships of Ancient Shallow-Marine and Slope Carbonates, Cambro-Ordovician Cow Head Group, Newfoundland: Implications for Fluid Flux. Journal of Sedimentary Petrology, 54(4): 1062-1080.

Sugiyama, K. and Anderson, O.R., 1997. Experimental and observational studies of radiolarian physiological ecology, 6. Effects of silicate-supplemented seawater on the longevity and weight gain of spongiose radiolarians Spongaster tetras and Dictyocoryne truncatum. Marine Micropaleontology, 29: 159-172.

215

Suhm, R.W., 1997. Simpson stratigraphy of the southern Midcontinent. In: K.S. Johnson (Editor), Simpson and Viola Groups in the southern Midcontinent. Oklahoma Geological Survey Circular. Oklahoma Geological Survey, pp. 3-38.

Sweet, W.C., 1954. Harding and Fremont Formations, Colorado. AAPG Bulletin, 38(2): 284-305.

Tabachnik, K.R., 1994. Distribution of recent Hexactinellida. In: R.W.M. van Soest, Rob W. M., van Kempen, Theo M. G., Braekman, Jean-Claude, Brink, Anita, Bianchi, Frans R., Vermuelen, Jan J. (Editors), Sponges in time and space. Biology, Chemistry, Paleontology. A. A. Balkema, Rotterdam, pp. 225-232.

Tanguay, L.H. and Friedman, G.M., 2001. Petrophysical characteristics and facies of carbonate reservoirs: The Red River Formation (Ordovician), Williston basin. AAPG Bulletin, 85(3): 491-523.

Tappan, H. and Loeblich Jr., A.R., 1973. Evolution of the oceanic plankton. Earth Sciences Review, 9: 207-240.

Textoris, D.A., 1968. Petrology of supratidal, intertidal, and shallow subtidal carbonates, , Middle Ordovician, New York, U.S.A. Internat. Geol. Cong., 23d, Prague, 1968, Proc., Sec. 8, Genesis and classification of sedimentary rocks: 227-248.

Thomas, W.A. and Astini, R.A., 1996. The Argentine Precordillera; a traveler from the Ouachita Embayment of North American Laurentia. Science, 273(5276): 752- 757.

Tobin, K., J., Walker, K.R., Steinhauff, D.M. and Mora, C.I., 1996. Fibrous calcite from the Ordovician of Tennessee: preservation of marine oxygen isotopic composition and its implications. Sedimentology, 43: 235-251.

Tolmacheva, T.J., Danelian, T. and Popov, L.E., 2001. Evidence for 15 m.y. of continuous deep-sea biogenic siliceous sedimentation in early Paleozoic oceans. Geology (Boulder), 29(8): 755-758.

Tomescu, I., Conley, J.A. and Kidder, D.L., 2003. Insight into spatio-temporal distribution of worldwide silurian cherts. Geological Society of America Abstracts with Programs, 35(6): 493.

Tomescu, I. and Kidder, D.L., 2002. Changing environments of chert accumulation through the Ordovician. geological Society of America Abstracts with Programs, 34: 554. 216

Tréguer, P, Nelson, David M., Van Bennekom, Aleido J., DeMaster, David J., Leynaert, Aude, Queguiner, Bernard., 1995. The silica balance in the world ocean: a reestimate. Science, 268: 375-379.

Trettin, H.P., 1970. Early Paleozoic Evolution of Northern Parts of Canadian Arctic Islands: Abstract. AAPG Bulletin, 54(12): 2509-2510.

Tribble, J.S., Arvidson, R., S., Lane, M.I. and Mackenzie, F.T., 1995. Crystal chemistry, and thermodynamic and kinetic properties of calcite, dolomite, apatite, and biogenic silica: applications to petrologic problems. Sedimentary Geology, 95: 11-37.

Tucker, M.E., 2001. Sedimentary petrology: an introduction to the origin of sedimentary rocks. Blackwell Scientific Publications, Oxford, Oxford, 252 pp.

Udden, J.A., 1921. Characteristics of Some Texas Sedimentary Rocks as Seen in Well Samples. AAPG Bulletin, 5(3): 373-385.

Van Houten, F.B. and Arthur, M.A., 1989. Temporal patterns among Phanerozoic oolitic ironstones and ocean anoxia. In: T.P. Young and W.E.G. Taylor (Editors), Phanerozoic ironstones. Geological Society Special Publications. The Geological Society, London, pp. 33-49. van Soest, R.W.M., 1994. Demosponge distribution patterns. In: R.W.M. van Soest, Rob W. M., van Kempen, Theo M. G., Braekman, Jean-Claude, Brink, Anita, Bianchi, Frans R., Vermuelen, Jan J. (Editors), Sponges in time and space. Biology, Chemistry, Paleontology. A. A. Balkema, Rotterdam, pp. 213-223.

Veevers, J.J., 2000. Analysis of Australia. In: J.J. Veevers (Editor), Billion-year earth history of Australia and neighbours in Gondwanaland. GEOMOC Press, Sydney, pp. 61-73.

Vermeij, G.J., 1995. Economics, volcanoes, and Phanerozoic revolutions. Paleobiology, 21(2): 125-152.

Vishnevskaya, V., 1997. Development of Paleozoic-Mesozoic Radiolaria in the northwestern Pacific Rim. Marine Micropaleontology, 30: 79-95.

Vishnevskaya, V.S. and Kostyuchenko, A.S., 2000. The evolution of radiolarian biodiversity. Paleontological Journal, 34(2): 124-130.

217

Vogt, P.R., 1989. Volcanogenic upwelling of anoxic, nutrient-rich water; a possible factor in carbonate-bank/reef demise and benthic faunal extinctions? Geological Society of America Bulletin, 101(10): 1225-1245.

Von Rad, U. and Rösch, H., 1974. Petrology and diagenesis of deepsea cherts from the central Atlantic. In: K.J. Hsu and H.C. Jenkins (Editors), Pelagic Sediments on Land and Under the Sea. International Assoc. Sedimentologists Spec. Publ., pp. 327-347.

Wainright, P.O., Hinkle, G., Sogin, M.L. and Stickel, S.K., 1993. Monophyletic Origins of the Metazoa: An Evolutionary Link with Fungi. Science, 260(5106): 340-342.

Webby, B.D., 1981. The Ordovician System in Australia, New Zealand and Antarctica. International Union of Geological Sciences, 6: 64 p.

Webby, B.D., 1984a. Early Phanerozoic distribution patterns of some major groups of sessil organisms, Palaeontology : proceedings of the 27th International Geological Congress, Moscow 4-14 August 1984 : volume 2. Utrecht, The Netherlands : VNU Science Press, Utrecht, pp. 222.

Webby, B.D., 1984b. Ordovician reefs and climate: a review. In: D. Bruton, L. (Editor), Aspects of the Ordovician system. Palaeontological contributions from the University of Oslo. Universitetsforlaget, Oslo, pp. 89-100.

Webby, B.D., 1995. Correlation of the Ordovician rocks of China : charts and explanatory notes. International Union of Geological Sciences publication ; no. 31. Trondheim, Norway : International Union of Geological Sciences, 103 pp.

Wengerd, S.A., 1948. Fernvale and Viola Limestones of South-Central Oklahoma. AAPG Bulletin, 32(12): 2183-2253.

White, R.D., 1986. Cambrian radiolaria from Utah. Journal of Paleontology, 60(3).

Wiebe, W.A.V., 1948. So-Called Maquoketa of Barber County, Kansas: Geologic Notes. AAPG Bulletin, 32(1): 126-128.

Wilde, P., 1989. Oceanography in the Ordovician. In: C.R. Barnes and S.H. Williams (Editors), Advances in Ordovician Geology. Geological Survey of Canada, pp. 283-298.

Wilde, P. and Berry, W.B.N., 1984. Destabilitzation of the oceanic density structure and its significance to marine "extinction" events. Palaeogeography, Palaeoclimatology, Palaeoecology, 48(2-4): 143-162. 218

Williams, L.A., Parks, G.A. and Crerar, D.A., 1985. Silica diagenesis, I. Solubility controls. Journal of Sedimentary Petrology, 55(3): 0301-0311.

Wilson, J.L., 1954. Ordovician stratigraphy in Marathon folded belt, West Texas. AAPG Bulletin, 38: 2455-2475.

Wise, S.W. and Weaver, F.M., 1974. Chertification of oceanic sediments. In: K.J. Hsü and H.C. Jenkyns (Editors), Pelagic sediments: on land and under sea. International Association of sedimentologists; Blackwell Science, Inc., pp. 301 - 326.

Won, M.-z. and Below, R., 1999. Cambrian Radiolaria from Georgina Basin, Queensland, Australia. Micropaleontology, 45(4): 325-363.

Wood, R.A., Dickson, J.A.D. and Grotzinger, J.P., 2002. Proterozoic modular biomineralized metazoan from the Mama Group, Namibia. Science, 296(5577): 2383-2386.

Woody, R.E., Gregg, J.M. and Koederitz, L.F., 1996. Effect of Texture on Petrophysical Properties of Dolomite: Evidence From the Cambrian-Ordovician of Southeastern Missouri. AAPG Bulletin, 80(1): 119-131.

Workman, L.E. and Bell, A.H., 1948. Deep Drilling and Deeper Oil Possibilities in Illinois. AAPG Bulletin, 32(11): 2041-2062.

Worsley, T.R. and Kidder, D.L., 1991. First-order coupling of paleogeography and CO2, with global surface temperature and its latitudinal contrast. Geology, 19: 1161- 1164.

Worsley, T.R., L., M.T., Fraticelli, C.M. and Scotese, C.R., 1994. Phanerozoic CO2 levels and global temperatures inferred from changing paleogeography. In: G.D. Klein (Editor), Pangea: paleoclimate, tectonics, and sedimentation during accretion, zenith, and breakup of a supercontinent. Geological Society of America Special Papers. Geological Society of America, Boulder, Colorado, pp. 57-73.

Xiaofeng, W. and Xiaohong, C., 1995. Ordovician paleogeography and biogeography in China. In: J.D. Cooper, M.L. Droser and S.C. Finney (Editors), Ordovician odyssey : short papers for the Seventh International Symposium on the Ordovician System. The Pacific Section Society for Sedimentary Geplogy, pp. 203-206.

Xiping, D., Knoll, A.H. and Lipps, J.H., 1997. Late Cambrian Radiolaria from Hunan, China. Journal of Paleontology, 71(5): 753-758.

219

Xu, C, Charles E. Mitchell, Zhang Yuan-dong, Stig M. Bergström, Donald Winston, Wang Zhi-hao, Jörg Maletz., 1995. Graptolite species succession across the base of the Undulograptus austrodentatus zone at Hunagnitang, south-central China, and its prospects fro use as a global stage stratotype. In: J.D. Cooper, M.L. Droser and S.C. Finney (Editors), Ordovician odyssey; short papers for the Seventh international symposium on the Ordovician System. Field Trip Guidebook - Pacific Section, Society of Economic Paleontologists and Mineralogists. Fullerton, pp. 498.

Yapp, C.J. and Poths, H., 1992. Ancient atmospheric CO2 pressure inferred from natural goethites. Nature, 355: 342-344.

Yiren, L. and Hanying, F., 1989. A candidate stratotype section of Ordovician Hanjiangan and Shikouan formations in China; Shuangjiakou Section at Qidong County, Hunan China; Part I. Dicengxue Zazhi = Journal of Stratigraphy, 13(3): 161-192.

Yuan, X. , Chen, Zhe, Hu, Jie, Parsley, Ronald L., Xiao, Shuhai, Zhou, Chuanming., 2002. Towering sponges in an Early Cambrian Lagerstaette; disparity between nonbilaterian and bilaterian epifaunal tierers at the Neoproterozoic-Cambrian transition. Geology, 30(4): 363-366.

Yu-Jing, W., 1988. Middle Ordovician radiolarians from the Pingliang Formation of Gansu Province, China Abstract. Geologica et Palaeontologica, 22(Special Issue First international conference on Radiolaria (EURORAD V); abstracts): 214.

Yu-Jing, W., 1991. On progress in the study of Paleozoic radiolarians in China. Acta Micropalaeontologica Sinica = Weiti Gushengwu Xuebao, 8(3): 237-251.

Zhang, X.-g. and Pratt, B.R., 1994. New and extraordinary Early Cambrian sponge spicule assemblage from China. Geology, 22(1): 43-46.

Zhiji, Z., Yunan, N. and Wenwei, Y., 1995. Outline of Ordovician paleogeography, Tarim, Northwest China. In: J.D. Cooper, M.L. Droser and S.C. Finney (Editors), Ordovician odyssey; short papers for the Seventh international symposium on the Ordovician System. Field Trip Guidebook - Pacific Section, Society of Economic Paleontologists and Mineralogists. Fullerton, pp. 498. 220

LOWER ORDOVICIAN APPENDIX

North America USA Depositional Chert No Formation Age Location Lat/Long Associated facies SSR* References environment type

1 Alsate Shale l, Ar Se Oklahoma, SW Texas, 34008'N/96056'W shelf B sh, LMS, SST Harlton (1966); Berry (1958) 2 Blakely Fm. l Eastern Ouachitas of Arkansas, 34031'N/94033'W undecided B sh, SST, slts Stone and Sterling (1962) 3 Chepultepec Fm. l Tennessee, Kentucky, W North Carolina, West Virginia, peritidal/lagoonal N, O Dol H. B. Renfro, Dan E. Feray (1989) Virginia, Maryland, Alabama, 36037'N/84046' W

4 Cool Creek Fm. l, Tr SW Oklahoma, 34025'N/97059'W peritidal/lagoonal N, B LMS, evap Ragland and Donovan (1991); Gao and Land (1991) shallow shelf 5 Cotter Fm. l E Kansas, W Missouri, Arkansas, 36047'N/94027'W peritidal/lagoonal N, O ? McCracken (1955) 6 Crystal Mountain Fm. l Arkansas, 34031'N/94004'W undecided B sh, SST Pitt et al. (1961) 7 Deepkill&Normanskill beds l, Ar E New York belt, 42 017'N/73048'W slope/basin B sh with grpt Rad Dietz and Holden (1966)

8 El Paso Fm. l New Mexico, Texas, Arizona 31058'N/105028'W peritidal/lagoonal N, B LMS, Dol Epis and Gilbert (1957), Lee and Friedman (1987)

9 Ellenburger Fm. l W Texas, Oklahoma 30025'N/103044'W peritidal/lagoonal N,O LMS, Dol SS Udden (1921); Cole, (1941) 10 Fort Ann Fm. l E New York, SW Vermont, 43012'N/73019'W peritidal/lagoonal N, B LMS Mazzullo and Friedman (1975) 11 Gailor Fm. l, Tr Central New York shallow shelf, N LMS, Dol Philips and Friedman (1995) subtidal-tidal 12 Gasconade Dolomite l S Missouri, W Tennessee, Kentucky, N Arkansas, peritidal/lagoonal N, O Dol Woody et al., (1996) 36045'N/89050'W 13 Goodwin Limestone l CS Nevada, California, Nevada, 37030'N/115048'W shelf N LMS Ross, (1964); Langenheim et al., (1962) 14 Great Meadows Fm. l E New York, SW Vermont, 43012'N/73019'W peritidal/lagoonal N, B LMS Mazzullo and Friedman (1975) 15 Gunter-Jordan Fm. l Illinois, 400N/88017'W peritidal/lagoonal N, O SST, Dol, sh Workman and Bell (1948); Buschbach (1964) 16 Jefferson City Fm l Kentucky , Oklahoma, Kansas, Missouri, 94033'N/36055'W peritidal/lagoonal N, O Dol, Sh SS Renfro and Feray (1989)

17 Kindblade Fm. l, Ar SW Oklahoma, 34025'N/98005'W peritidal/lagoonal N, B LMS SS Ragland and Donovan (1991); Latham (1968); Harlton (1963) 18 Kingsport Fm. l W North Carolina, E Tennessee, E Kentucky, West Virginia, peritidal/lagoonal N, O LMS, Dol Renfro and Feray (1989), Perry et al. (1979) W Virginia, Maryland, 36037'N/84046'W 19 Lambs Chapel Dol l S Ohio, 39024'N/83047'W peritidal/lagoonal N, O Dol Dolly and Busch (1972) 20 Longview dolomite or Nittany l Alabama, SW Virginia, Georgia, E Kentucky , Tennessee peritidal/lagoonal N, B, O LMS, Dol Bennison (1975); Folk and Pittman (1971) dolomite 34044'N/85018'W shallow shelf 221

Depositional Chert No Formation Age Location Lat/Long Associated facies SSR* References environment type

21 Manitou Fm l central Colorado, 38 044'N/105027'W shelf N, B Dol SS Gerhard (1974) 22 Mascot Fm l W Tennessee, Kentucky 36034'N/88052' W peritidal/lagoonal N LMS, Dol Renfro and Feray (1989) 23 Mazarn shale l Arkansas, 34031'N/94004'W slope/basin B grpt black sh Pitt et al. (1961); Finney (1997) 24 McKenzie Hill Fm. l W Oklahoma, 35013'N/95000'W undecided N LMS, Dol Latham (1968) 25 Oneota dolomite l Wisconsin, SE Minnesota, N Iowa, S Wisconsin, peritidal/lagoonal N Dol Fisher and Barratt (1985); Prokopovich (1953) 43014'N/91010'W 26 Ontelaunee Fm. l W Pennsylvania, 40024'N/75038'W undecided N Dol, LMS Hobson (1957) 27 Phi Kappa Fm. l central Idaho, 43 053'N/114034'W slope/basin B sh with grpt; Q Churkin (1963) 28 Powell-Smithville Fm. l E Kansas, W Missouri, Arkansas, 36047'N/94027'W peritidal/lagoonal N, O Q, sh McCracken (1955) 29 Rickenbach Fm. l W Pennsylvania, 40024'N/75038'W undecided N, B Dol Hobson (1957) 30 Roubidoux Fm. l Missouri, S Kansas, Oklahoma, 37018'N/94033'W peritidal/lagoonal N,O LMS, SST McCracken (1955); Overstreet et al. (2003) 31 Shakopee Fm. l Wisconsin, SE Minnesota, NE Iowa, 42 058'N/90043'W peritidal/lagoonal O, undet LMS, Dol Fisher and Barratt (1985); Smith et al., (1993)

32 Siliceous Lime l Oklahoma, Spavinaw, 34044'N/97016'W shelf N LMS, Dol Edson (1927); Gallardo and Blackwell (1999) 33 Stonehenge Fm. l North Carolina, Tennessee, Kentucky, West Virginia, peritidal/lagoonal N LMS, Dol Renfro and E. Feray (1989), Moshier (1986) Virginia, Maryland, Pennsylvania 36016'N/82012' W shallow shelf 34 Tribes Hill Fm. l Central New York 42050'N/75036' W shallow shelf N LMS, Dol Philips and Friedman (1995) 35 Turkey Mountain lime l NE Oklahoma, Tulsa 36017'N/95045'W undecided N LMS Ruedemann and Redmon (1927) 36 West-Spring Creek Fm l, Ar SW Oklahoma, 34020'N/97043'W shelf N,B LMS Ragland and Donovan (1990); Smith et al. (1993) 37 Windfall Fm. l, Tr Nevada, 39031'N/115058'W shelf undet lime mudst, sh SS, Rad Kozur, et al (1996); Bird (1989)

Canada 38 Dildo sequence l NE Newfoundland, 49 052'N/55060'W slope/basin B argillite, volc Kay (1967) 39 ? l Gaspe Peninsula, 48046'N/65017'W slope/basin B sh, cgl, SST Hein and Parrish (1989) 40 ?Little Port Complex l W Newfoundland, the Bay of Islands, 46 031'N/67041'W slope/basin B ophiolitic complex Rad Blome et al. (1995b); Aitchinson et al. (1998)

41 lower part of Cow Head group l W Newfoundland, 50 0N/580W slope/basin N, B LMS, sh with grpt SS, Rad Suchecki and Hubert (1984)

42 Neruokpuk Fm. l Yukon Territory, Barn Mountains, 68 054'N/139033'W undecided B LMS, volc Churkin, Jr. (1975)

43 Snook's Arm group, Wild Bight l NE Newfoundland , W Notre Dame Bay, 49 054'N/57010'W slope/basin N, B LMS, sh, volc., SST Hein and Parrish (1989) Gps. 222

Depositional Chert No Formation Age Location Lat/Long Associated facies SSR* References environment type South America Bolivia 44 Cieneguillas Argillites l, Tr Tarija region,South Bolivia,21030'S/64040'W slope/basin B black sh Erdtmann et al. (1995)

Australia and Oceania Australia 45 Howqua shale l E Victoria, 370S/1470E slope/basin undet Webby (1981) 46 Oolloo Limestone l N territory, Daly River Basin, 150S/1320E shelf undet LMS Webby (1981)

New Zealand 47 Anthill Black shale, Aorangi l NW South Island, 41 051'S/172001'E slope/basin B black sh with grpt, Rad Hein and Parrish (1989) Mine Fm. SST

Europe England 48 ? l S Scotland, 57 004'N/4013'W slope/basin B volc, black sh Rad Grunau (1965); Hubert (1966) 49 Durness Limestone, Ballantrae l Ballantrae, Scotland, 58 015'N/4020'W slope/basin; N, B ophiolitic complex SS, Rad Grunau (1965); Aitchison (1998); Armstrong et al. (1999) Complex peritidal/lagoonal 50 Raven Gill Fm. l, Ar Southern Uplands, Scotland, 55 0N/50W slope/basin B oph. complex Rad Floyd (1996)

Norway 51 Valhallfonna Fm. l, Ar Spitzbergen = Svalbard, 78045'N/15032'E shelf undet LMS Rad Blome et al. (1995); Renz (1990)

Poland 52 Holy Cross Mount. l Holy Cross Mountains, 50 040'N/210E undecided undet ? Fischer (1984)

Asia Ural Area 53 ? l Urals, 57039'N/65059'E slope/basin B sh Rad Hein and Parrish (1989) Kazakhstan 54 Burubaital Fm. l 46053'N/66018'E slope/basin B oph. complex SS, Rad Blome et al. (1995b)

Uzbeksiatn 55 Besopan Suite l Kyzylkum desert, 42030'N/650E slope/basin undet turbiditic seq. Rad Drew et al. (1996) 223

Depositional Chert No Formation Age Location Lat/Long Associated facies SSR* References environment type Korea 56 Mungok Fm. l, Tr Yongwol, Taebaeksan basin, 37011'N/128028'E shelf, middle and B LMS, sh Kim and Choi (2000) outer ramp

China 57 Beilongmiao Fm. l, Tr Xichuan, Yangtze platform, SC China, 32 030'N/1100E shelf B LMS, Dol Webby (1995)

58 Fenxiang Fm. l Yangtze George area, western Hubei, 30030N/109 030'E shelf undet LMS SS Liu et al. (1997)

59 Guanyintai Fm. l, Tr central China, 32 0N/1200E shelf B Dol Webby (1995) 60 Honghuayuan Fm. l Yangtze George area, western Hubei, 30030N/109 030'E shelf undet LMS SS Liu et al. (1997)

61 Liangchiashan Fm. l Liangchiashan near Shimengzhai, 400N/1200E shelf N LMS, Dol Webby (1995) 62 Woboer Fm. l, Ar Hangwula Mountains, Alxa, Nei Monggol, 41030'N/1040E shelf B dark argill, slate Webby (1995)

63 Zhifangzhuang Fm. l Shangdong, North China Platf., 36 030'N/1170E shelf undet Dol Webby (1995) Others 64 ? l,m Sonora, Mexico; 110 052'W/28036'N slope/basin B black sh, SST, LMS Finney and Bartolini (1993)

65 Garden City Fm. l,m SE Idaho, SW Wyoming, USA, 110003'/42031'N peritidal/lagoonal, N, B LMS Berry (1962) shelf 66 Heituao Fm. l,m Queerqueke area, Tarim block, NW China, 40 054N/88020E slope/basin B black sh Rad Zhiyi et al. (1995)

67 Karmberg Limestone l,m Australia, Tasmania, 146028'E/42040'S shelf N LMS Webby et al. (1985)

MIDDLE ORDOVICIAN

North America USA 1 Barrel Spring Fm. m Mazourka Canyon, California, 36 052'N/117037'W middle-outer shelf N SST, sh, LMS Langenheim et al. (1956); Stokstad (1995) 2 Blackford Fm. m Virginia, 36059'N/82049'W undecided N, B LMS, sh Kay (1948) 3 Chazy limestone m Ohio, C Pennsylvania, 40039'N/78014'W peritidal/lagoonal N LMS, sh Dolly and Busch (1972); Conley (1977) shelf 4 Fort Pena Fm. m SW Texas, 30024'N/103023'W peritidal/lagoonal, N, B, O sh, clcr, cgl SS Wilson (1954); McBride (1969) slope/basin 5 Hardy Creek limestone m SW Virginia, 36050'N/81048'W undecided N LMS, argill. LMS Harris (1958) 224

Depositional Chert No Formation Age Location Lat/Long Associated facies SSR* References environment type

6 horizon no.2 & 3 of the Vinini- m Nevada, E California, 38013'N/118035'W slope/basin B LMS, sh, volc Rad D.E. Feray et al. (1968); Murchey et al. (1983) Valmy Fm. 7 Lincolnshire limestone m West Virginia, Tennessee, 36050'N/82049'W shelf, subtidal N, B LMS SS Kay (1948); Bauert (1994) 8 Mazourka formation m California, Badger Flat, 36052'N/117037'W shelf, subtidal N, B LMS Langenheim et al. (1956); Sloan (1976) 9 Pinesburg Strata m E Kentucky, W Virginia 36040'N/82042'W undecided N Dol, LMS Renfro and Feray (1989); Bennison (1989)

Canada 10 ?Bay of Islands Complex m W Newfoundland, the Bay of Islands; 46 031'N/67041'W slope/basin B ophiolitic complex Rad Blome et al (1995b)

11 Cobbs Arm sequence m NE Newfoundland, 46 031N/67041W slope/basin B Volc, LMS, argill sh Kay (1967) with grpt 12 Red shale, Cow Head Gr. m W Newfoundland, 50 0N/580W slope/basin N, B LMS, sh with grpt SS, Rad Suchecki and Hubert (1984); Coniglio (1987)

Australia and Oceania Australia 13 Numeralla formation, m S New South Wales, 38 004'S/144043'E slope/basin B sh with grpt, Rad Dietz and Holden (1966); Fergusson and Fanning (2002) Adaminaby beds turbiditic seq. 14 Temperance Fm. m SC New South Wales, 36 030'S/149030'E slope/basin undet Volc., SST, sh Rad Webby (1981) 15 Wahringa LMS m New South Wales, 38 006'S/147020E undecided B LMS Lyons and Percival (2002)

Europe England 16 Kirkton Fm., Crawfrod Gr. m Southern Uplands,Scotland, 55 0N/50W slope/basin N, B mudst Rad Floyd (1996); Danelianand Clarkson (1998)

Germany 17 Sylt ?? m Sylt, N. Germany, 54054'N/8022'E undecided undet ? Schallreuter (1989)

Norway 18 ? m Trondheim, 63052'N/11042'E slope/basin B volcanics Rad Hein and Parrish (1989)

Asia Ural Area 19 ? m Ural area, 57039'N/65059'E slope/basin B ? Rad Blome et al (1995b);

Kazakhstan 20 Burubaital Fm. m Kazakhstan, 46053'N/66018'E slope/basin B ophiolitic complex SS, Rad Blome et al (1995b); Tolmacheva et al. (2001) 225

Depositional Chert No Formation Age Location Lat/Long Associated facies SSR* References environment type

21 Bestamaksky LMS m eastern Kazakhstan, 450N/750E undecided undet sh with grpt, LMS Rad Renz (1990) 22 Tekturmas Fm, Urtynzhal Gr m Dzhungaria-Balkhash zone, Tekturmas Mt. 46015'N/660E undecided B slts, breccia, volc Rad Nikitin et al. (1986)

China 23 Klimuli Fm. m Tongxin, Ningxia, Zhuozishan, 390N/106050'E slope/basin undet LMS, sh with grpt Rad Yu-jing (1991) 24 Lanmutan Fm. m Sandu, Yangzte platf, 26 0N/1080E shelf undet argill LMS Webby (1995)

Others 25 Clements Fm. m,u Australia, CW New South Wales, 35 0S/1470E slope/basin undet slate, cgl, Q Webby (1981) 26 Hoskins Chert, Jindalee gr m,u New South Wales, Australia, 33 052S/148 0E slope/basin B ? Lyons and Percival (2002) 27 Hulo Fm. m, NW and NE Jiangxi, China, slope/basin undet black sh, LMS Rad Xu et al. (1995) 28 Mugincoble chert m,C New South Wales, Australia, 38 006S/147 020E undecided B ? Lyons and Percival (2002) 29 Ottawa rocks m,u N Mississippi, USA, 89015'W/34035'N shallow shelf N LMS Mellen (1974) 30 Pingliang Fm. m,C Gansu, China, 35 035'N/106031'E slope/basin undet LMS, grpt sh Rad Yu-Jing (1988) 31 Postolonnec Fm. m, u the Armorican Massif (W France), Crozon Peninsula, SW shelf N mudst, or clayey slts Loi and Dabard (2002) Sardinia, 48015N/4030W 32 Road River Fm. m,u N Yukon & Alaska, Canada, 137 020'W/66054'N slope/basin N, B grpt. sh, argill LMS, Jackson and Lenz (1962) Dol, SST 33 Shoal Arm Fm. m,u NC Newfoundland, Canada slope/basin B argil, black sh, flysh Rad Bruchert et al. (1994) 34 Sofala volcanics m, C Australia, CW New South Wales, 33 0S/1500E slope/basin undet volc SS, Rad Webby et al. (1981) 35 Womble Shale m, C Arkansas, USA, 94033'W/34031'N slope/basin B black sh, LMS Stone and Sterling (1962) 36 Yoshiki Fm. m,u Central Japan ? B volc Rad Igo et al. (1980)

UPPER ORDOVICIAN

North America USA 1 Aleman Fm. Ash New Mexico, W Texas, SC Mexico, 31058'N/105025'W subtidal, shelf N, B LMS SS Pope and Steffen (2002); Brimberry (1984)

2 Beckett Fm. C SE Missouri, 37012'N/90033'W shelf N LMS, cgl, calc Larson(1951); Craig (1993) 3 Bigfork Fm. C Oklahoma, 34039'N/94026'W slope/basin B sh Campbell and Suneson (1989) 4 Black River limestone C Virginia, Tennessee, 43002'N/84032'W peritidal/lagoonal O, undet LMS, Dol Cohee (1948); Textoris (1968) shelf 5 Brannon limestone C SC Kentucky, 36058'N/82025'W shelf N LMS SS McFarlan and White (1948); Fisher (1968) 6 Carters limestone C W Tennessee- Kentucky, 36034'N/88052'W subtidal, shelf N LMS Renfro and Feray (1989); Holland and Patzkowsky (1997) 226

Depositional Chert No Formation Age Location Lat/Long Associated facies SSR* References environment type

7 Curdsville limestone C SC Kentucky, 36058'N/82025'W shelf N LMS McFarlan and White (1948); Fisher (1968) 8 Deceit Fm. u Alaska, Kotzebue Sound, 65 030'N/1660W slope/basin B LMS, grpt sh Harris et al.(1995); Ryherd and Paris (1987) 9 Deicke K-bentonite beds*** C NW Georgia, S Minnesota, NW Iowa, E Missouri, Kentucky, - B ash beds Huff and Kolata (1990) C Tennessee 10 Duzel Formation u N California, Sierra Nevada, 42055'N/122013'W slope/basin B LMS Feray et al.(1968); Condie and Snansieng (1971) 11 Ely Springs Dol Ash SC Nevada, 37011'N/115033'W shelf N, B LMS, Dol Burchfiel (1964); Miller (1977) 12 Fernvale Fm. C,Ash SC Oklahoma, 34044'N/97009'W subtidal, shelf N LMS Wengerd (1948); Denison (1997) 13 Fish Haven Fm. u Idaho-Utah; 42 002'N/112046'W shelf N Dol SS Beus (1968); Churkin (1962) 14 Fremont Fm. u W Colorado, E Utah, 38 049'N/109002'W shelf N, B LMS Sweet (1954); Gerhard (1967) 15 Galena Fm. C Iowa, 41056'N/93027'W shelf N, B LMS, Dol Agnew (1955); Finney (1997) 16 Guttenberg limestone C Iowa, Missouri, 40031'N/93023'W shelf N LMS, sh Agnew (1955); Finney (1997) 17 Hager Fm. C SE Missouri, 37012'N/90033'W shelf N, B LMS Larson (1951); Craig (1993) 18 Hanson Creek Fm. Ash C Nevada, 38054'N/116046'W shelf, slope/basin N, B LMS SS, Rad Blome et al. (1965b); Finney et al (1999) 19 Joachim dolomite C SE Missouri, 37012'N/90033'W shelf N, B Dol Larson(1951); Craig (1993) 20 Keel Fm. Ash Oklahoma, Texas, 35045'N/100038'W shelf N LMS, Dol Kopaska-Merkel and Friedman (1989); Qing Sun (1995)

21 Kimmswick Fm. C SE Missouri, 37012'N/90033'W shelf N LMS Larson(1951); Craig (1993) 22 Livengood Dome Fm. u EC Alaska, 64042'N/143026'W slope/basin B sh, tuff, LMS, Hein and Parrish (1989); Harris et al. (1995) siltstone 23 Macy Fm. C SE Missouri, 37012'N/90033'W shelf N, B LMS Larson(1951); Craig (1993) 24 Maquoketa Fm. Ash S Kansas, Illinois, EC Iowa, 340N/97016'W outer shelf N, B Dol, argill rock SS Wiebe (1948); Grammer (1983) 25 Maravillas chert C,Ash W Texas, 30027'N/103023'W slope/basin N, B LMS, sh SS, Rad King (1926) 26 Martinsburg Fm. Ash Virginia, New York, 390N/790W slope/basin B sh, turbidites Rad Mcbride (1962); Carroll (1959) 27 McGregor limestone C E Iowa, 41037'N/91001'W shelf N LMS, Dol Byers and Stasko (1978); Agnew (1955) 28 Millbrig K-bentonite bed*** C NW Georgia, S Minnesota, NW Iowa, E Missouri, Kentucky, - B ash beds Huff and Kolata (1990) C Tennessee 29 Munsungun Fm. C NW Maine, 46030'N/680W shelf, slope/basin B sh, turbidites SS, Rad Pollock (1987) 30 Murfreesboro-Pondspring Fm. C W Tennessee- Kentucky, 36044'N/88052'W subtidal, shelf N LMS Renfro and Feray (1989); Holland and Patzkowsky (1997)

31 Phi Kappa Fm. C,Ash NE Nevada, 41059'N/115035'W slope/basin B grpt sh, LMS, Dol Rad Paull and Paull (1981); Cone et al. (2002) 32 Polk Creek Shale u Arkansas, 34031'N/94033'W slope/basin undet sh Stone and Sterling (1962) 33 Red River Fm. Ash North Dakota, Montana, 49 0'N/101031'W shelf N, B LMS, dol Tanguay and Friedman (2001); Andrichuk (1959)

34 Saturday Mountain Fm. C,Ash Idaho, 43044'N/113023'W shelf undet grpt sh, LMS Churkin (1962) 35 Sympson shale C SC Oklahoma, 34044'N/97016'W peritidal/lagoonal N, B sh, LMS, SST Edson (1927); Suhm (1995) 36 Trenton Fm. C NW Ohio, 41010'N/81018'W offshore, shelf N LMS, Dol Stieglitz (1975); Gregg and Sibley (1984) 37 Tyrone limestone C SC Kentucky, 36058'N/82025'W peritidal/lagoonal O, N, B LMS, Bentonite McFarlan and White (1948); Bauert (1994) shelf 227

Depositional Chert No Formation Age Location Lat/Long Associated facies SSR* References environment type

38 Valmont dolomite of the Ash Texas, SC New Mexico, 31060'N/105006'W shelf N, B LMS, dol P'An (1982) Fusselman Fm. 39 Vinini-Valmy Fm. C,Ash Nevada, E California, 38013'N/118035'W slope/basin B LMS, sh, volc. Rad Feray et al. (1968); Noble and Finney (1999) 40 VIOLA zone III C S Kansas, SC Oklahoma, 98005'/37014'N shelf, slope/basin N, B Dol Wiebe (1948); Grammer (1983)

Canada 41 ?Bay of Islands Complex C W Newfoundland, the Bay of Islands, 46 031'N/67041'W slope/basin B ophiolitic complex Rad Blome et al. (1965b)

42 Advance Fm. C British Columbia, Advance Mt, 550N/1250W slope/basin N LMS, Dol, sh Norford et al. (1995) 43 Allen Bay carbonate rocks u Cornwallis and Griffith Islands, Queen Elizabeth Islands, shelf, slope/basin N, B Dol, LMS Sodero and Hobson (1979) 75065'N/94059'W 44 Blue Star Fm. u NC Newfoundland, 49 0N/55030'W slope/basin B volc-sed rocks Kusky and Kidd (1996) 45 Halysites-Bearing Beds u Alberta, British Columbia, 51 039'N/116040'W undecided B Dol Severson (1950) 46 Hazen Formation C,Ash Arctic Canada in northeastern Ellesmere, 79 037'N/81052'W slope/basin B grpt sh Rad Gentzis et al., (1996); Trettin (1970)

47 Stony Mountain Fm. Ash SW Manitoba, Gretna, 49 050'N/100014'W shelf N sh, LMS Andrichuk (1959); Porter and Fuller (1959)

Australia and Oceania Australia 48 Ballast beds u SC New South Wales, 32 045'S/145048'E slope/basin undet slts, slates,grpt Rad Webby (1981); Iwata et al. (1995) 49 Birkenburg beds C, Ash SC New South Wales, 35 027'S/149049'E slope/basin undet slts, slates,grpt Rad Webby (1981) 50 Bogong Creek beds C SC New South Wales, 36 0S/1480E slope/basin undet slts, slates,grpt Webby (1981) 51 Bubalahla Fm. C Rockley - Taralga area, 34030S/149 040'E slope/basin undet grpt black sh Webby (1981); Lyons and Percival(2002) 52 Bumballa Fm. C SE Australia, Cooma-Mandurama, 36.5 004'S/148043'E slope/basin B grpt sh Rad Dietz and Holden (1966);Fergusson and Fanning (2002)

53 Cotton beds Ash Gunning - Blandarea, 33 0S/148020'E slope/basin undet LMS, grpt sh Webby (1981); Lyons and Percival(2002) 54 Jerrawa beds C, Ash capital territory, Yass, 35 0S/1490E slope/basin B grpt sh Webby (1981) 55 Malongulli Fm. C, Ash Cliefden caves, 33050'S/148025'E slope/basin B LMS, tuff, grpt sh SS, Rad Rigby and Webby (1988) 56 Nine Mile Volcanics C Kiandra, 360S/1490E slope/basin B basalt, tuff, sh Webby (1981) 57 Uralba beds C Attunga, 30055'S/150050'E undecided undet mudst, LMS Rad Webby (1981) 58 Wagonga Fm. C Narooma, 36014'S/150004'E slope/basin B volc, grpt black sh Webby (1981) 59 Warbisco shales, Mallacoota C SE Australia,E Victoria, -38 004'S/144043'E slope/basin B grpt sh Rad Dietz and Holden (1966); Webby (1981) beds

Europe England 60 ? - a flysh seq. C S Scotland, 57 004'N/4013'W slope/basin B black sh, mudst, Rad Grunau (1965); Hubert (1966) 228

Depositional Chert No Formation Age Location Lat/Long Associated facies SSR* References environment type

61 Currarie Fm. C Scotland, 55 0N/50W slope/basin B basalt, grpt Rad Armstrong et al.(1998); Floyd (1996) 62 Glenkiln Fm. C Scotland, 55 0N/50W slope/basin undet mudst, grpt sh Rad Armstrong et al.(1998) 63 Hartfell Shale Fm C Scotland, 55 0N/50W slope/basin undet mudst, grpt sh Rad Floyd (1996) 64 Marchburn Fm. C Scotland, 55 0N/50W slope/basin B basalt, grpt Armstrong et al.(1998); Floyd (1996)

Italy 65 Portixeddu Fm. u SW Sardinia, 39030'N/8030'E shelf N muds, slts SS Loi and Dabard (2002)

France 66 Kermeur Fm. u Armorican Massif, 48015'N/4030'W inne-outer shelf N muds, slts Loi and Dabard (2002)

Asia Ural Area 67 ? C Urals, 57039'N/65059'E slope/basin B ? Rad Blome et al. (1995b)

Kazakhstan 68 Ul'Kuntassk Fm. C,Ash Chuillisk Mountains, 46 053'N/66018'E slope/basin B LMS Rad Blome et al. (1995b); Nazarov and Ormiston (1993)

69 ?Chu-Illisk Mt & rocks Ash Chu-Illisk Mounatins, 44 0N/750E undecided undet LMS Rad Renz (1990) ofPredchingiz 70 Kuvsky Fm. Ash Predchingiz, 46015'N/660E undecided B volc, LMS, silts Rad Nazarov and Ormiston (1993)

China 71 Chengbu Fm. C Shuangjiakou, 26049'N/112007'E slope/basin B silic black sh Yiren and Hanying (1989) 72 Gangmusang Fm Ash Xainza, Gangmusang, 310N/88030'E shelf N LMS, Dol Webby (1995) 73 Shuangjiakou Fm. C Shuangjiakou, 26049'N/112007'E slope/basin B silic black sh Yiren and Hanying (1989) 74 Wufeng Fm. Ash Yangzte platf, 300N/107030'E slope/basin undet silic grpt sh, LMS Rad Yu-jing (1991)