z o in X z I ^ I i y i ^ j‘/>N0linIlJ.SN|ZNVIN0SHllWSWS3 I HVB 8 nZLI B RAR I ES^SMITHSONIAN INSTIT (/) _ Z \ ^ Z ^ ay Z c/J Jfe, 2 A \ o 2 5 2* y/*' - V l^L! B RAR I ES^SMITHSONIAN^SNSTITUTION NOlinilJLSNI^NVINOSHIIWS^SB I 9' ^SS\ § ./ v A m /^kwi&s. c w r /0\ t f\ S % rrs .... w _ NOliniUSNrNVINOSHilWS^Sa I HVH aifllB RAR 1 ES^SMiTHSONIAN INST!* tel 2 \L g m z m ___ ^ ^ 1 L. I 8 RAR I ESVnSMITHSONIAN_!NSTITUTION' NOlifUllSNi “NVINOSH1I INS S3 I H z •» ' s> X£v osy^X > N“/N0liniliSN|ZNVIN0SHlllNs'/>S3 I aVdSn^LIB RAR I ES^SMITHSONIAN JNSTI (/> — ”3* \ 2 JML, Z o w 2 5 2 r 5 v>' n"JLIBRARIES SMITHSONIAN INSTITUTION NOlinillSNI NVIN0SH1IINS S3 1 i 1 V z r- z r* ._, Z 03 DN *v 2 s s) > 1 5/ 33 P&mM&r&y n \iM4Mw — " vo; m Ng.D-vr 2; w''' ^ ^ W NOIlfliliSNI- NVINOSHillNS^SS I aVaan~UBRAR I ES ^SMITHSONIAN JNST m b KAK I to v>(Vl( I noUi^jMIN IliO I I I U I Iwli IMUIXI (MWUHWOnAI^O » ova a 8 CO _ Z CO Z W OlinillSM NVINOSHilWS S3IBV88n LIBRARIES SMITHSONIAN INSTITUTIO “ - m ^ m . _ _ co' £ co X 5 JOliniUSNI NV1N0SH1IINS S3lUVd8l1 LIBRARIES SMITHSONIAN INSTITUTIC Z £ 5 ,v<’' £ 2 ,v. •1 I vyy-t I 1 s \aszy 1 1 Xf l .1 B R AR I ES^SMITHSONIAN INSTITUTION NOliniliSNI NVINOSHilWS^SS I « VB 9 s *' \^V > ' 2 X^osv^vx >’ JOlinillSNI^NVINOSHlIWS^SB I HV8 8 !l\l B RAR S ES^SMITHSONIAN JNSTITUTIO 32ESSS. w V w v> xfSESx “ w h mv,4 ~ h ~ Wm£%. h . ,^c _ '' R 2 *J Z _ JBRARIES SMITHSONIAN INSTITUTION NOIifUIiSNI NVIN0SH1MS S3I8YB8 2 z r- z lOIlfUIiSNI NVINOSHIIINS^SS I HV8 8 I1~L I B RAR I ES ^SMITHSONIAN INSTITUTE z CO z g z vX 5! J&' , \ INTRODUCTION Sacred bundles were a part of the religious and ceremonial regalia of the Arapaho Indian religion. They usually contained objects empowered with spiritual and ritual significance. One such bundle is in the Arapaho collection of Carnegie Museum (Number 3179A/147). This collection, acquired by the Museum from the Fred Harvey Company, was assembled by Dr. George A. Dorsey, presumably in 1901 or 1902, when he was study- ing the Arapaho Indians of Oklahoma. The purpose of this paper is to describe the bundle. This will be accomplished by (1) general remarks on sacred bundles and their place in the religion of the Plains Indians; (2) a complete quotation from the catalogue of the collection compiled by Dorsey; and (3) a description of the bundle and its contents in order as they were unwrapped by the writer. In a concluding section the bundle will be compared with one described by A. L. Kroeber, the only other known description of an Arapaho bundle. SACRED BUNDLES IN THE RELIGION OF THE PLAINS INDIANS Sacred bundles were peculiar to the religion of many of the Plains Indians of the United States, as well as to some of the tribes living along the Mississippi River and the western Great Lakes. These bundles usually contained a variety of objects believed to be endowed with spiritual powers and to have originated in a supernatural experience. As such, the bundles were regarded as extremely sacred (unlike the paraphernalia of the age societies which was more or less secular). They were opened only at certain religious ceremonies, and in a hallowed atmosphere.* The Arapaho Indians did not have an extreme development of the bundle complex as did, for example, the Pawnee and the Blackfoot Indi- ans. Further, it appears that bundles were not regarded by the Arapaho to be as sacred as their sacred pipe and wheel (Kroeber, p. 309-311). Neverthe- less, the bundles were of considerable importance in Arapaho religion, and the opening of them involved much ritual. These bundles were kept in the tribe and passed from one generation to another in accordance with the “keeper-owner” concept of ownership. Among some tribes, for example the Blackfoot, the bundles could be sold or transferred by one owner to another (Wissler, 1912, p. 272-278). Exact Arapaho customs as regards own- ership are not known; Kroeber reported the bundle which he described to ^Present day Plains Indians still regard these bundles with sincere respect and religious awe. Mr. Hassrick states that he has seen Indians refuse to be present when a bundle has been opened in a museum. Their feeling and respect for these sacred objects are very real. 11 Issued June 20, 1957. 12 Annals of Carnegie Museum vol. 35 have been acquired through inheritance (Kroeber, p. 310). This was prob- ably the normal method of passing on bundles, with the owner serving as a keeper or custodian of a bundle for the other tribal members. DORSEY’S INFORMATION ON THE BUNDLE An account of this bundle was given by Dorsey with the collection as it came to Carnegie Museum; apparently he simply recorded the informa- tion given by his Indian informant. Some of this information is not too clear, and there are some statements which appear inaccurate according to the physical facts of the bundle. It is possible that the informant was speaking of sacred bundles in general, or did not remember the contents of this bundle too clearly. Dorsey’s description of the bundles is quoted in full below, with in- formation in brackets supplied by the present writer. WAHSAWK (Bear’s Bag), or SACRED BUNDLE. The whole oblong bag is a chain of mountains, with a man sitting on it, watching constantly, and at the same time feeding the inhabitants of the earth. Any person, especially the owner of it, stores away in it goods, provisions, valuable articles, for the future. Inside, next to the designs, thus making it complete, are found good dried buffalo meat, elk meat, thick pieces of fat, pemmican, mixed berries, strouding, calico, coffee, sugar, etc. Since these Indians are poor and can not afford to put away such things, this Wahsawk has no contents, excepting a Dog-soldier’s paint string. There were some spruce needles in the bag, but an old man bought them some- time ago, for his future use. This sacred bag, as also others, was made after the buffalow [sic.] cow, which afterwards pitied him, appeared to the hunter several times [in a vision.] This man and his wife had one sore-backed pony. The camp moved quite a distance, which caused them to camp near the creek, which had plenty of timber. During their camp a change of life took place with them. On their return to the tribe, the man erected an Old Men’s Lodge and his wife erected the Buffalo Women’s Lodge. Thus the Wahsawk bags were made with society rules and regulations. After the men have reached the old men’s lodge, they have walked to the top of the mountain, i.e., attained old age. At the bottom of the “Image” on each side, are bear’s claws, typifying the person begging the “god” for mercy and continuous blessing. The foundations of the image are a leg bone of the bear, several bladder bags, a serpent's body in hide and several bunches of human hair. Inside of the image there is found buffalo wallow hair. The sticks are used for the braces of a dragging-cart for it [the bundle]. The feather, loose in the bag, is from a whip-poor-will. In all the designs the beads and porcupine quills are used to signify persons and animals. The sticks are from a bush called “basanechi” (big garter-snake), not used for a dragging cart; but they are all utilized for various purposes — canes, pokers, and inside braces for beds. The feather of the whip-poor-will dramatizes praying all night, or continuous worship, without fatigue. The porcupine loops denote tipis and furniture. On these designs are represented the Four Old Men. Both of the coverings are from “yellow” calves, — meaning bright days, prosperous years, blessings upon the people. The calf’s hide next to the image is from an extraordinary [unusual, perhaps abnormal] calf, thus making it charitable, generous and humane to the image and more natural to the people. The buffalo wallow hairs represent society lodges; the image, a creator and the earth. Dorsey’s information will be amplified and perhaps clarified by the following description and by comparing this bundle with the one described by Kroeber. 1957 Stoller: A Sacred Bundle of the Arapaho Indians 13 A DESCRIPTION OF THE SACRED BUNDLE The part$ of the bundle are described and illustrated in the order in which they occurred when the bundle was unwrapped. The various cover- ings or bags, and all of the thongs and ties had been thoroughly impreg- nated with red ocher. This had also rubbed off on most of the contents of the bags giving everything a reddish-brown color and perhaps con- tributing to the pungent odor of the bundle. The outer layers of the bundle. The second bag was loosely wrapped in a piece of figured calico material and tied with a rawhide thong. The material was a floral print in blue, black, and yellow. This in turn was wrapped in a second piece of calico which partly protruded from the bundle and was quite dirty. This material was a brown, blue, and white plaid. The outer covering was a rather roughly dressed deer or elk skin. It was much worn and weathered. Few traces of red paint remained on the exterior. This bag is shown in Fig. 1. It was roughly constructed, apparently from as whole skin. The skin; was slit where a seam was necessary and then sewn with large stitches of sinew. One pair of ties over the top remained, and there was evidence of another pair. A long, thick thong tied the whole together. Complete dimensions of the bag were: length, about 26 1/2 inches; width, 14 inches; and height, 10 inches. The second bag. The second bag contained the innermost bag. When this most sacred part of the bundle was removed there was found at the bottom of the second bag a large undecorated “envelope,” made of roughly dressed hide, partially filled with dried wild plums ( Prunus sp.). The envelope is shown in Fig. 2. It was rectangular, with a triangular flap fastened by two short thongs. There were two sets, each of two thongs, on the upper back of the bag, apparently to hold a suspension cord. The envelope was 15 inches long, 12i/4 inches wide at the bottom, and IO14 at the top. The second bag is shown in Fig. 3. It was lined with canvas or duck around the sides. Stippling in the illustration indicates this canvas lining. The sewing was done with large stitches of sinew. The bag itself was made of well dressed buffalo hide (probably from a young animal) and was quite worn. It had been torn and patched many times, especially around the top where the ties were. Two sets of two thongs each were placed below the edge, about five inches apart, on each of the long sides. Apparently these ties were fastened first, and then the long thongs, one at each end, were tied over them pulling the large ends over the opening of the bag for further protection of the contents. These thongs were then tied around the bag. A looped strap for carrying or suspending the bag was fastened to the back. The ends of the bag were decorated with a beaded strip of black and white “pony” beads. The complete design is shown, reconstructed, at the left in Fig. 3. Dorsey’s explanation stated that the beads, like the quills. 14 Annals of Carnegie Museum vol. 35 Fig. 1. Exterior of sacred bundle 1957 Stoller: A Sacred Bundle of the Arapaho Indians 15 Fig. 2. Envelope containing dried food 16 Annals of Carnegie Museum vol. 35 1957 Stoller: A Sacred Bundle of the Arapaho Indians 17 signified persons or animals. The design may represent the Four Old Men of Arapaho mythology. This bag was approximately 25 inches long, 12 14 inches wide, and 9 14 inches high when filled with its contents. The inner bag and its contents. The contents of this bundle were the most important of the whole bundle; that is, the remainder consisted mostly of coverings. Although the entire contents of the innermost bundle were completely removed for examination, the bundle or bag itself— folded and laced together with heavy thongs— was not completely unfolded. Fig. 4 shows its appearance and gives some idea of the way in which it was constructed. It would seem that the piece of buffalo calf skin was first doubled, and then at either end the separate folds were folded inward and laced with heavy thongs. The method of folding the skin explains why the fur appears both on the inside and outside (shown in the illus- tration by hatching). Since the bag was made of buffalo skin it was quite bulky and fairly heavy. It measured 21 inches in length and nine inches in width (at its widest part) and was approximately six inches high. The thong, shown partly under the bag in the figure, was loosely wrapped around it. The length of the looped thong was approximately 31 inches, while it was five-eighths of an inch wide. In the bottom of the bag, partially buried by the folds, was a polished stick placed lengthwise; a second such stick was loose inside the bag. It is shown in Fig. 5, at A. The ends are carved smooth, and the whole is polished and has a slight sheen. One end was slightly larger than the other. The stick illustrated was 21 inches long, and had a diameter of one-half inch at the largest end. Dorsey reported that the sticks were from a bush called “basanechi” (big garter-snake). He first noted that the sticks were used for a dragging cart for the bundle (a kind of travois), but, further on, his information was contradictory for he said that they were not used for a dragging cart but for canes, pokers, bed braces, etc. In Kroeber’s description of a sacred bag he mentioned “four small sticks of unequal length, which are leaned up together like tentpoles, and which represent a tent” (Kroeber, p. 311). This may be some form of an altar. He also referred to a fifth stick which was kept inside the case and “represents the gift of a horse.” This bag had a total of five sticks, two of which were incorporated in the base of the “image” and one in the vertical part of it (See below). They were several inches shorter than the two described above. A knotted thong was lying in a heap in the bottom of the bag. A por- tion of it is shown at B in Fig. 5; it was 53 inches long. Like all the other articles of hide, it had been covered with a reddish paint, presumably red ocher. It was a dog-soldier’s paint string, according to Dorsey, and contained 89 knots. The Dog Lodge is the fourth of the Arapaho age societies (open to men around fifty years of age), and, according to Kroeber, painting was an important part of the initiation ceremony (Kroeber, p. 199-200). 18 Annals of Carnegie Museum vol. 35 Fig. 4. The inner bag 1957 Stoller: A Sacred Bundle of the Arapaho Indians 19 Fig. 5. A, One of four sticks of various lengths. B, Knotted thong. C, Piece of skin 20 Annals of Carnegie Museum vol. 35 Every morning each dancer, accompanied by his wife, repairs to his grandfather. His grandfather has a string with a number of knots tied in it, representing the number of times he was painted when he went through the ceremony. This number is often about forty, but sometimes runs as high as ninety. Too great a number is thought to be undesirable. Every time the grandfather paints the dancer he unties one knot in his string. At the same time the younger man makes a knot in a string which he brings. At the end of the painting the grandfather has thus untied all the knots in his string, and his grandson’s string now contains the same number, which, when he in turn becomes the grandfather of a later dancer, will be untied in the same way. A fragile, dried piece of skin was folded in the bottom of the bag. It is shown at C in Fig. 5, and was identified as pigskin or peccary skin. Bits and pieces of fibers and a seed were embedded in the fur of the bag. The remnants of two feathers— mostly the quills— were also in the bottom. A third feather was found between the two folds of one side of the bag. This apparently was the feather referred to by Dorsey as being from a whip-poor-will and represented continuous worship. The feather is sketched at B in Fig. 6. It was b]/2 inches long. It was reddish in color as a result of some of the bag’s paint rubbing on it. Together with the items discussed above were found three bladder bags. These bags were in poor repair. The bottom bag contained no opening except for a few torn holes. It was empty but may once have contained the spruce needles referred to by Dorsey. This bag was about 12 inches long. One of the other bags is shown at A in Fig. 6. It was filled with porcupine quills, mostly white ones, but a few were dyed yellow or red. The bag was split lengthwise and laced with sinew at both ends. The third bag contained mostly yellow quills with a few white ones; it was made like the one just described but was laced with commercial thread. The two bags were of equal length, 10 i/2 inches. The quills had not been flattened as was the custom of the Arapaho women in preparing them for use in decoration. In the top of this bag, loosely tied in by the aforementioned thong, was the “image.” It is shown in Fig. 7 (with the base slightly foreshortened). The “image” was not entirely disassembled in this examination, since it was so thoroughly bound and fastened together. Dorsey mentioned a serpent’s skin, but this was not in evidence in the “image”; possibly the skin in question was the one described above and found loose in the bag. The foundation, or cross-bar of this inverted T-shaped object, consisted of several things all bound together with thongs wound around them. There were two polished sticks, one of which protruded at either end. These sticks were surrounded by bladder bags. Another stiff element ran lengthwise of the base, but it was wrapped with quill work and was not otherwise visible. It might have been the bear’s leg bone referred to by Dorsey; it was about the same size as the sticks, with diameters not ex- ceeding one-half inch. Large and small loops of white porcupine quills were incorporated in the base. At one end was a doubled piece of hide, partly protruding, the edges of which were stitched with quills. Wisps of hair (Dorsey said it was human hair and this seems likely) stuck out in various places. 1957 Stoller: A Sacred Bundle of the Arapaho Indians 21 Fig. 6. A, Bladder bag with porcupine quills. B, One of three feathers; this one separate from others Annals of Carnegie Museum vol. 35 1957 Stoller: A Sacred Bundle of the Arapaho Indians 23 The vertical part of the “image” was encased in a cone-shaped, thick, and well dressed piece of hide laced up one side. It was held, none too securely, to the foundation by a few thongs which passed around and were incorporated in the base. At the bottom of this piece were several vertical elements wedged between the horizontal ones of the base. Among these were three fairly short objects wrapped in quills and narrow thongs. These were the bear claws referred to by Dorsey except that instead of two there were three, and they were in the middle and not at the ends of the “image.” The fifth polished stick stood inside the case; it was the shortest of all, being only 7 14 inches long. The remainder of the contents of this vertical section consisted only of buffalo hair stuffed inside. The dimensions of the “image” were, length, 15 inches; height, 10i/2 inches; height of the vertical piece only, approximately 8 inches; approximate diameter of the bottom on the vertical piece, 4i/2 inches. Dorsey gave the symbolism of the various parts of the “image.” The bear claws represented a person begging the “god” for mercy and blessing. The quills were supposed to represent persons and animals, while the quilled loops stood for tipis and furniture; designs on them were supposed to represent the “Four Old Men,” but since they were devoid of design perhaps this remark referred to the beadwork on the outer cover. The buffalo hairs represented society lodges (or degrees), while the “image” itself represented a creator and the earth. A COMPARISON WITH THE BUNDLE DESCRIBED BY KROEBER Krober’s remarks on a sacred bundle are here quoted in full (Kroeber, p. 310-311). There are a number of sacred bags in the tribe, and one was obtained, through inheritance, by a young woman shortly before the summer of 1899. In Plate LXIII the bag is shown unpacked, and its contents in the position in which they are set in the tent. The upright cylindrical leather object is a person, apparently the spirit of the bag. Food and gifts were given to the bag on the occasions on which it was opened. The food, or part of it, was then kept in the bag. The object of these gifts was to secure prosperity and an abundance of food to the giver. The disposition of the contents of the bag during the ceremony is as follows: — The owner of the bag sits at the rear of the tent. In the middle is a small fire of coals. Directly before the owner is the outer bag, now holding only the various gifts of food contained in small bags and pieces of cloth. The outer bag itself is still covered, to a large extent, with the calico with which it is ordinarily kept wrapped, and which presumably has been given to it. The smaller longi- tudinal bag is removed from the inside of the large bag, and put in front of it, nearer the fire. In front of this are placed grizzly-bear claws. To the right of these claws is a small bag of incense called “man,” and to the left a similar bag called “woman.” Immediately in front of the grizzly-bear claws are four small sticks of unequal length, which are leaned up together like tentpoles, and which represent a tent. In front of these sticks, and not far from the fire, is the cylindri- cal leather case representing the “owner” of the bag. By the side of this object and the four sticks, another stick, ordinarily kept inside of the cylindrical case, is laid on the ground. This represents the gift of a horse. This is all the exact information Kroeber gave on the subject of sacred bundles. He did not specify on what occasions the bundle was opened. 24 Annals of Carnegie Museum vol. 35 whose tent was used, who opened the bundle, and who could or was expected to give it gifts. There is little reason to doubt that the bundle described by Kroeber and the one described in this paper are of the same type. Some details differ, however. The bundle in Carnegie Museum had three cases or bags instead of two; this bundle contained no bags of incense unless they were embedded in the foundation of the “image" (which was not taken apart); the “image” in Kroeber’s bundle did not have a foundation or base attached as this one did; the feather, bladder cases of quills, and knotted thong were unique to this bundle. Perhaps they had been gifts to it. Kroeber’s description made clear some of the remarks in Dorsey’s cata- logue notes. For instance, Kroeber explained some of the contents of the bundle as being gifts to it, apparently as a form of supplication or offering. The “image” represented a person and the spirit of the bag. This is a less esoteric explanation than Dorsey’s that it was “a creator and the earth.” Kroeber’s remarks suggested something of the ceremony involved in opening such a bundle. Apparently only the owner could open it; there was an established order for removing the contents, and they were placed in pre-determined positions. Each object had symbolic significance. Dorsey’s information was concerned mostly with this symbolism. Dorsey’s references to the Four Old Men were apparently from the lore surrounding the Sun Dance; neither Kroeber’s account of the bundle nor his discussions of Arapaho religion and ceremonies contained such an allusion. The Sun Dance was one of the biggest and most important cere- monies in Arapaho life. Dorsey identified the Four Old Men as “the gods of the four world quarters . . . the Four Old Men, are considered as ever- present, ever-watching sentinels, always alert to guard the people from harm and injury” (Dorsey, p. 14). It is possible that this bundle may have been used in the Sun Dance ceremonies. ACKNOWLEDGMENTS The author is grateful to Carnegie Museum, and especially to James L. Swauger, Assistant Director, for permission to describe this bundle. She would also like to acknowledge the assistance of Mr. Swauger and Miss Kay Reichert in reading and criticizing the manuscript. Royal Hassrick, of the Denver Art Museum, also read the paper and gave it the benefit of his wide knowledge of the Plains Indians. Dr. J. K. Doutt, Curator of Mammals, identified all the skins in the bundle, and Dr. L. K. Henry and Mrs. Dorothy L. Pearth of the Section of Plants identified the plums in the second bag. J957 Stoller: A Sacred Bundle of the Arapaho Indians 25 REFERENCES Dorsey, George A. 1903. The Arapaho sun dance; the ceremony of the offerings lodge. Field Columbian Museum Anthropological Series, v. 4, 228 p. Chicago. Kroeber, A. L. 1907. The Arapaho. Bulletin of the American Museum of Natural History, v. 18, pt. 4, p. 1-229, 279-454. New York. Lowie, Robert H. 1916. Plains Indian age-societies; historical and comparative sum- mary. In “Societies of the Plains Indians,” edited by Clark Wissler. American Museum of Natural History Anthropo- logical Papers, v. 11, p. 877-984. New York. Wissler, Clark 1912. Ceremonial bundles of the Blackfoot Indians. American Mu- seum of Natural History Anthropological Papers, v. 7, pt. 2, p. 65-289. New York. 1920. North American Indians of the Plains. American Museum of Natural History Handbook Series, no. 1, 172 p. New York. m.-u t5 HP\A ART. 3. REMARKS ON THE BEHAVIOR OF THE SQUIRREL TREEFROG, HYLA SQUIRELLA By Coleman J. Goin and Olive B. Goin Carnegie Museum and University of Florida INTRODUCTION Several years ago we became} interested in the day-by-day activities of the local treefrogs and began keeping detailed notes on individuals observed during a continuing study of our back-yard fauna. Although much of this information has been published elsewhere (Goin and Goin, 1953; Goin, 1955) we have a residue of data on the non-breeding behavior of these frogs which we believe should be made available. The present paper deals with the habits and habitat of the Squirrel Treefrog, Hyla squirella Sonnini and Latreille. The observations on which this paper is based were made in our back yard near Gainesville, Alachua County, Florida. The yard is a plot nearly rectangular in shape and approximately a third of an acre in extent, bordering on a typical North Florida mesic hammock. The most conspicu ous and abundant trees are water oaks ( Quercus nigra), but in addition there are black gums ( Nyssa sylvatica biflora), sweetgums (Liquidambar styraciflua), hickories ( Carya sp.), blue beeches ( Carpinus caroliniana), several live oaks ( Quercus- virginiana), one small holly ( Ilex sp.), a magnolia sapling ( Magnolia grandiflora), and a large loblolly pine ( Pinus taeda). The undergrowth in the yard has been cleared out but the leaf mold has been left undisturbed. Bordering the yard on three sides, however, is a dense thicket of seedling oaks, palmettos ( Sabal minor) and shrubs and vines. Observations in previous years on Hyla squirella had indicated that these frogs are present in considerable numbers in the yard during the fall, winter and spring months. Further, we had observed that they are noc- turnal, stirring with the approach of darkness and settling down just after daylight. They show a strong tendency to return time after time to the same spot to rest during the day. We also noted that when the temperature drops below a certain level the frogs do not stir from their hiding places at night. It was in order to determine, if possible, the temperature at which activity ceases that we initiated an intensive study of the movements of these frogs in the fall of 1953. Late in August we set six stakes about four feet high across the eastern end of the lot. On top of each stake we inverted an empty tin can. This row of stakes with their cans was about fifty feet east of the house, parallel to and just at the edge of the woods bordering the yard. From October 22, 1953, until May 31, 1954, these cans were examined morning and evening for the presence of Hylas, as were the undergrowth bordering the yard and various other habitats, both natural and edificarian, in which we knew treefrogs were apt, to occur. 27 JWG 1 2 m Issued July 2, 1957. 28 Annals of Carnegie Museum vol. 35 FALL RETURN In Florida Hyla squirella breeds in the spring and summer, utilizing temporary ponds and flooded ditches as well as more permanent bodies of water. Not all of the frogs leave the area around our house at this time, for we hear the rain call throughout the summer months. This rain call should not be confused with the call given in the breeding choruses; in warm, humid weather it is uttered sporadically by resting individuals at any time during the day. Apparently the frogs that remain in the yard during the summer move high up in the tops of the oak trees for we almost never see them. The summer of 1953 was one of the wettest ever recorded for northern Florida. The yard was flooded much of the time during late August and early September and although we observed no breeding activity, we did record treefrogs more frequently than normal for this time of year. Usually the reappearance of adult frogs in the hammock occurs quite abruptly toward the end of September. In the fall, young squirella begin to invade the hammock in increasing numbers. We never saw more than one immature a day from the time the cans were put up until October 30 when we saw three. On October 31 we recorded 10. The young frogs seem to move into the hammock in waves. For the next 13 days we saw an average of four a day, then on November 14 we counted 15. There was apparently another wave on November 21, when we recorded 10, and another on November 24 when 17 were seen. This was the greatest number observed on any one day during the course of the study. On December 8, we recorded 16. Thereafter the fluctuations in number from day to day seemed more to reflect variations in temperature than to indicate fresh invasion. NUMBERS IN YARD Since the preceding summer had been an unusually favorable one for breeding, the number of young squirella in the yard during the course of thel study probably represents a high point in the population cycle for the area. It is difficult to form an accurate estimate of the actual number of frogs resident in the yard. It seems probable that not all of those which were noted when one of the “waves” moved in, found suitable niches and remained in the yard for the winter. The average number of young seen per day for each month was as follows: October (last week), 2.38; November, 5.97; December, 7.65; January, 4.81; February, 3.40; March, 3.29; April, 2.23; May, 2.43. This decrease from midwinter to late spring was not due simply to the growth of the individual frogs to a size at which they were no longer recorded as immatures. Even in April and May the young of the previous summer are recognizable as not full grown. Although we madei no measure- ments, our observations would tend to agree with those of Wright (1932, p. 322-323) that the one year olds are smaller than the two year old (and older?) ones. He lists 13 mm. as transformation size, 19-23 mm. as one year old and 26 mm. -f- for twoi year olds. 1957 Goin and Goin: Behavior of the Squirrel Treefrog 29 The decline in number of young recorded in the spring was undoubtedly due in part to mortality. We observed no actual instances of predation, but on one occasion a Rough Green Snake ( Opheodrys aestivus) was found on two consecutive days coiled on the branches of a young water oak in the thicket just back of can no. 3. Several times Yellow Rat Snakes ( Elaphe obsoleta quadrivittata), which we have previously reported (Goin; 1955, p. 101) preying on Hylas, were found in the vicinity. At 9:15 on the evening of February 4, a small Yellow Rat Snake was seen on a bar of the rack for the fuel oil tank against the back of the house. The temperature was then 59°. Two adult H. squirella and one H. cinerea had been on the bar during the day. Now one squirella remained on a leg of the rack below the Elaphe. At 11:00 P.M. the snake was still there and the frog jumped to the ground. The next morning, at 9:00 o’clock with the temperature at 46°, the Elaphe wasi coiled on the bar, a squirella was sleeping on a brace bar about an inch below the snake, and an Anolis, which had not been present the night before, was resting about 6 inches farther along on the same bar as the snake. At 10:30 the temperature had warmed up to 56° and the snake was beginning to stir a little. By 11:45 it was crawling slowly along the bar. The Anolis and squirella were still quiescent. Half an hour later, the snake was where the lizard had been, the lizard was gone, and the frog still slept on the brace bar below. The next time the rack wa4 inspected the snake had, left. Adult squirella were much less numerous than were the young. We never saw more than three on any one day. The total number of records for each month were as follows: October (last week only), 7; November, 22; December, 10; January, 28; February, 44; March, 26; April, 0; May, 10. Many of these records represent the same frog seen day after day. DEPARTURE IN SPRING Carr (1940, p. 61) records breeding choruses of squirella in Florida from April 2 to August 20. No adult squirella were observed in the yard during April. The 10 records listed for May represent what was probably a single individual (possibly two). In marked contrast, the young were nearly as numerous in April (61 records) and May (66 records) as they were in March (78 records), and were still present when the study was terminated in June. This suggests that yearling squirella do not leave the winter habitat to move to the breeding ponds with the adults. During the humid rainy season in the summer they move higher up into the oak trees where they escape observation. Probably it is these yearlings that are responsible for the rain calls heard in the non-breeding habitats in the summer months. SELECTION OF HABITAT-YOUNG Table 1 gives the sites in the study area selected by immature squirella for daytime resting places from October 22, 1953 to May 31, 1954. Of the edificarian situations, the “rack” is the metal stand for the fuel oil tank which is located against the back of the house. The “pipe” is an iron pipe an inch in diameter, with one end angled upward to a height of 50 Annals of Carnegie Museum vol. 35 fcj ft* 5 O' EC! w D h go 5 c/5 5< . w >i C/5 « k . o Q g HQ u w w w J pa w a So $o pa s §g Q Q O >< C* < 120 27 42 O o> 150 26 89 in 550 52 156 rO 00 00 338 00 00 00 ro — — ** CD — 99 O 66 ro — 30 CO 40 O 00 ro OJ CD ro 00 o> m CO 37 34 Is- CD Z < O CAN 2 CAN 3 CAN 4 JO z < Q z «£ O RACK PIPE OTHER TOTAL EDiFSCARIAN NATURAL OAKS HOLLY MAGNOLIA in U.I O H- 1— UJ 2 _J i OTHERS TOTAL NATURAL TOTAL 1957 Goin and Coin: Behavior of the Squirrel Treefrog 31 about 15 inches, lying on the ground behind can no. 3. Under natural situations, the records for the oaks and palmettos represent a number of stations in the edge of the thicket surrounding the yard. The holly, on the other hand, is a single small tree, standing in a rather isolated position in the yard about five feet west of can no. 2. The magnolia is an isolated sapling about five feet tall. A number of these records repre- sent a single frog returning day after day to the same resting place. On the other hand, not all the records from one niche represent a single frog; for example, at least four frogs are known to have been involved in the records for cans no. 1 and no. 6. It is apparent that there is a shift during the winter in the niches chosen by young squirella. When they first move into the yard, they do not select diurnal resting places that will serve throughout the winter. They rather seem to occupy the first available site they find and are much less apt to return to it the next morning. Apparently neither the young oaks nor the palmettos provide niches sufficiently sheltered to be utilized throughout the winter. All of the 52 records for young squirella on oaks and all but four of the 81 records on palmettos were made during October, November and December. Only the holly, and to a lesser extent the magnolia, were utilized consistently until late spring when some of the leaves begin to drop from these broad-leaved evergreens. The suitability of the holly as a winter niche is, we believe, a reflection of the structure of the leaves. They are leathery and tend to curl over at the edges. A resting squirella clings to the under side of one of these leaves, fre- quently choosing one that is lying on top of another leaf so that the frog can find better protection on the holly than on oak or palmetto leaves. Edificarian situations were preferred to the majority of natural ones as permanent niches. During November, when the young frogs were just moving into the yard, 115 were found in natural situations and 64 in edificarian ones. In December there was a marked reversal with 62 frogs found in natural situations and 176 in edificarian. During February and March both types of habitats were utilized in approximately equal num- bers, largely because of the availability of the holly. During May only edificarian sites were occupied. One noticeable phenomenon was the failure of the frogs to utilize can no. 4 although they were observed in the other cans a total of 423 times. We suspected at first that there might be something about the can itself that repelled the frogs. On December 11 the can was exchanged for a; fresh one and the old can placed on the outside kitchen window sill. The next morning) a squirella was resting on the can on the sill. We suspect now that the absence of frogs in no. 4 was due to its position in the middle of the row. There was simply a much greater probability that a frog, as it moved along the edge of the wood, would reach a can at the end of the line and stop than that it would continue until it came to no. 4. 32 vol. 35 Annals of Carnegie Museum SELECTION OF HABITAT-ADULTS Table 2 shows the niches selected by adult squirella as diurnal resting places. In contrast to the young, adult squirella are much more inclined to select edificarian situations as soon as they arrive in the yard. Of 147 records, only five are from natural situations, four from the holly in October and November and one from a palmetto in November. In further contrast to the young, however, the adults did not utilize the cans to any extent (four records during October and November). The most favored site was the metal rack for the fuel oil tank. We recorded adult squirella from this rack 94 times from October through March. The other 44 records were all from situations around or close to the windows of the house— a jar on the kitchen window sill, a crack between the screen and window frame, the back porch, etc., so that all but nine of the records for adults were of frogs on or about the house. Probably this preference of the adults for niches close to the* house is correlated with availability of food supply. The lighted windows attract insects in the evenings and the insects in turn attract the frogs. When these frogs seek daytime resting places, they simply go to the first sheltered niches they find. In marked contrast, only 31 of the 888 observations on immatures were made on or about the house. This suggests that there may be a difference in diet between the adults and young. Perhaps the latter feed more on ground-dwelling forms such as ants rather than on the flying insects which are attracted to the lighted windows. REACTIONS TO TEMPERATURE Under appropriate weather conditions, squirella leaves its place of diurnal abode about dark and returns again about dawn the following morning. It was in order to determine the effect of temperature on these nightly wanderings that we began to make a check for the presence of Hylas after dark as well as in the mornings, recording whether or not the frog had stirred from its niche, and noting the temperature at the time the observa- tion was made. We also recorded the low temperature for the diel period and the temperature at which the niches were inspected the fol- lowing morning. The temperature readings were taken from a Taylor maximum-minimum thermometer fastened about five feet from the ground on an oak tree in the middle of the yard. A total of 974 observations is summarized by means of the bar diagrams in Fig. 1. The discrepancy between this figure and the totals in the habitat tables results from the fact that occasionally a frog was disturbed when the niches were inspected in the morning or for some other reason left during the day. 1 In each of these diagrams the vertical line represents the mean, the horizontal line represents the range, and the bar represents one standard deviation on each side of the mean to indicate dispersion. It can be seen that, even though there is overlap, there is a pronounced tendency for the frogs to remain in on cold nights and to forage on warm nights. It can further be seen that the young and adults react in essentially the same manner. TABLE 2. SUMMARY OF DIURNAL NICHES SELECTED BY ADULT HYLA SQUIRELLA DURING THE NON-BREEDING SEASON 1957 33 Goin and Goin: Behavior of the Squirrel Treefrog 94 44 142 o o o 24 CM 26 rO rO 44 GO o 28 CO CM o CM ; C". © © CM © CM CO © CM c/5 -rf © r-3 © © © © © © 0) (MC)oqC50-HOoqi')-.; i.-; < tO^OOtOOOt'iftOOOMOO ^ t o i i^. /*/> — - _f -H (M (M fa O 2 fa o©O)®a)co-H^>^O0ai!oaooo©oo(DN!O nfa POh h Wfa w fa og c o Hctf < fa c4 < > OJ©OC'I©(M©0i©i>»©iri©»r3©£'^© 3 © £ 4h 73 b0 03 c 73 © a a a © a © LH o <4H O Lh l2 © 1 4- 7s Lh o 03 73 © © © ccS O h bo S3 fa .2? ’5 & Is o H '2 o3 u 5h CD S3 HH *x o3 s 5h C/5 o bb S3 : © CM co co f-H © © © 1-H <-H C/5 Of 05 05 CM pH © © GO G 2 c^S ^ G 'G GO oo © CO o CM CO © © CM © o © © © © © © © © © « a QJ a + + + + + + + + + + + + + c*1 £j CM © 05 o CO 00 05 I>» 05 © Tt< QO oo © co CO © © CO © © rn CU o CO Tf © Th 05 © CM CM CM CM CO CM CM co GO CM CM CM CM © © OO CM (j , , . © © © © © © © p4 © © © © © © © © © © © © © © © © © © V t+H fcg o .2 > +1 +1 +1 fl +1 +1 +1 il fl +1 fl +1 +1 +1 +1 fl +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 QJ QO CO 00 05 CO © CM © 00 o © o co © CM t"* CO CO © r-H © © o a cq CM QO QO CM © th 00 05 00 © © © © © CO 00 © CM 05 U > CO ^ QO © © »-4 CM CM CM CM CM CO CO CO TJH CM CM CO CM © © CO CM pj b 53 S CO Tj a « +1 ^ s Jh QJ -Q a .g s £ QJ to a £ n CO r-H © CM © © © © Tfl © CO CO pH CM CM CM pH r-H CM CM pH r-H pH pH CM CM CO © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © CO © © © © © © © © © +1 +1 +1 +1 +1 +1 +1 +1 fl +1 +1 +i fl fl fl fl fl fl fl fl fl fl fl fl fl fl © © c^« © CO CO © © © © CM CO CO OO 00 © CO © © CM © CO CO QO QO 00 © © CO CO — H CM CM CM i— J CM CM >-H l—H CM r-H © tJh CO CM !>• 00 00 CO 00 © © CM CM CO CO GO GO CM CM CM CM CM CM © £— CM © © CM CM © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © © CO © © © © © © © +1 +1 +1 +1 +1 +1 +1 fl fl +1 +1 +1 +1 fl fl fl f! fl fl fl fl fl fl fl fl fi © © © © GO t"- OO © © © © © OO CM © © CM © pH TjH © CO © CM 05 CO © CO 00 © © CM © 05 05 00 00 oo t"- CM CM CM QO © G © QO © CO CO © CO CO CM CM © © G c4 GO CO © © QO QO CO OO QO QO ~ £ CM CM 1—1 1—1 CM CM r— 1 © © © © CM OO CO CM CM © © © © CM CM © CO GO CO QO Th OO QO p4 p4 © QO l>- © CM 05 <-H r-H CO CM i—l CM CM CM CM 1—1 f-H QO QO 00 GO CO CO i— t r—t © QO co CO tf QO co © CM CM CO CM © © 00 00 G G © © CM 05 CO CM © CO © © GO © CM © © CM CM CO CM CM rJ, •-4 © © c4 CM CM © © © © CM CM QO QO r-H S— 1 CM CM i— i *— • r-H ^OOt^QQ^W^OO^O+^O+fOOM^Ot^lOt^oOt^OOt^OO 0105 © © *0 'B bb to G QJ T3 P3 QJ Jh 40 Jh rt G ^ ’V £ £ G bO Cu s 5 a a s s u< J |o mb ^ *53 «« ,0 5» QJ 2 I CL '£■} V G QJ I— I o * 50 Annals of Carnegie Museum vol. 35 Bias in external measurements In order to demonstrate the extent of variation due to bias, samples of Blarina brevicauda were taken from six localities in the northwestern one-sixth of Pennsylvania. Measurements of total length, length of tail, length of ear and of hind foot were all taken in the field (using a steel millimeter rule) with the exception of those from East Springfield, Erie Co., Pa. In this case the collector used calipers, which were then read against a steel rule. The following list groups the localities examined according to possible variations in measuring technique. Locality External Skull Number Month measurements measurements of of taken by taken by specimens collection A. East Springfield, Erie Co. W. W. Goodpaster B. State game lands 154, Erie Co. C. Benson Swamp, 5 mi. E. of Columbus, Warren }»C. L. Gifford Co. D. 1.5 mi. N. of Pittsfield, Warren Co. G. 2 mi. E. of Mars, Butler Co. H. 1.5 mi. N. of Darlington Beaver Co. }■ J.J. Christian ■J. E. Guilday 77 Oct.-Nov. 18 Oct. 19 Aug. 29 Sept.-Oct. 23 Sept.-Nov. 25 July-Aug. In the graphs. Fig. 4, statistics of three external measurements are pre- sented for each of the six localities. The shrews from each locality were aged by the method outlined above and only those which fell into age group 2 were used. These samples, therefore, represent shrews of as nearly as possible a comparable age group. Both males and females were included in the samples. In all the graphs (total length, length of tail, and hind foot) the means for the East Springfield group (locality A) are seen to be significantly lower than the means of localities B, C and D, although the populations are only some fifty miles apart. In each case locality A appears signifi- cantly smaller. Localities B, C and D show no significant differences between their respective means, but in length of tail and length of hind foot appear to be significantly greater again than localities G and H. These differences between the means, especially in length of tail (graph 2) and length of hind foot (graph 3) are associated with variation in meas- uring techniques. The measurements from locality A were taken with calipers and the means of the external measurements were, as a rule, consistently lower than those of the other five localities examined, which were measured directly against a steel rule. Localities B, C and D were measured by a second collector, while localities G and H were measured by a third collector. Skull measurements taken by a single individual Guilday: Variation in Blarina from Pennsylvania 51 Fig. 4. Comparison of population samples of Blarina brevicauda from northwestern one-sixth of Pennsylvania illustrating bias due to collectors’ techniques 52 Annals of Carnegie Museum vol. S5 and involving only one technique do not show this pattern. This makes it appear likely that the observed differences are the result of bias intro- duced by differing measuring techniques. In this case the collection was accompanied by sufficient data to detect the bias; however, in most cases the magnitude of this significant source of error would not be known. In view of this, external measurements of small mammals should be used with care in comparing populations. Cranial measurements The cranial measurements were much more constant than were the external measurements, the coefficients of variation in the former averaging much lower. Several measurements, however, are subject to differences which might affect comparisons of samples, and these should be kept in mind when they are used. Highly significant (probability 0.01) sexual differences occurred in the following measurements— total length of skull, length of upper tooth row, and occipital height (Table 2). Significant or highly significant age dif- ferences occurred in the following measurements— total length of skull, cranial breadth, maxillary breadth, rostral breadth, length of upper molari- form row, width of foramen magnum and occipital height (Table 1). Even though these differences are significant, they are slight and might not affect conclusions concerning geographic variation among large samples. However, with the small samples generally available, such differences should be recognized in order to avoid erroneous interpretations as to the taxonomic status of the populations represented. GEOGRAPHIC VARIATION In this survey, 1193 skulls of Blarina brevicauda, from 26 areas in Penn- sylvania (Fig. 5) were aged and measured, using the techniques outlined above. All immature, senile, and obviously abnormal individuals were discarded. In order to have the samples as large as possible the sexes were combined in all cases except weight. The differences are slight and the sex ratio in the samples not markedly biased. Weight is treated for each sex separately because it varies seasonally in response to the repro- ductive cycle. In southeastern Pennsylvania four of the areas (designated as V, Fig. 5) were lumped into one sample because of the small number of specimens from each. Likewise, two areas (T, Fig. 5) in northeastern Pennsylvania were combined to make one sample. The resulting data are shown by graphs (Fig. 6-14) and in Tables 3-15. All the graphs except that of weight agree in a common geographic placing of localities. The graph for weight deviates from this plan, in that the localities are placed in a seasonal sequence rather than a geo- graphical one. Each graph is split into three divisions. Localities A through J are all located in western Pennsylvania; localities K through P are lo- cated in central Pennsylvania, and localities Q through V in the eastern part of the state. As one progresses from left to right within each of the three divisions one travels from north to south; that is, locality A is in extreme northwestern Pennsylvania, locality J, extreme southwest, etc. Guilday: Variation in Blarina from Pennsylvania 55 ■§22l^3'S lN.§zg<3«| o -« .2 2 oi 25-s a a s- 3 s "S o “ • « ~ U s S ^ w* -5^ ,oOM ^ 2 W G <<0 o C_3 •« S 3.- *3 • u ^ r,3 -S ^ 0 ** O ^ fj* Q* > O * &.S §2 o ^ § css ^.2 S 8 1 g .a:50! jj ,§|£ l^oeil^ ^ r~ • 2* • ^ (U fr, O "* - W 8-u*“ T ^ S d 5 u -~ ' . . on O »® R-v "tj • iJ *”1 <£ O S £l >^5 * D v h > c ^ £11 » . § °._--§ .i’s S1*« as® g § = oo a d 3 ,-es 8 2 2 3 a ® oA’S I ?•§ a f k § 2 m b ; S £\® a s Ki . o 5 ^ O ■« v . *r3 12 « S O ?> . . *> *> _ £ ? bd ^ ^o-a t? ss'sari sli 1 «o|-_-ii|’S "u ssaBaSsBjf «a0 «g|02 w* £ O o S»3 o 1 ^ o ^ s;c]^ ^ r4 ho o v ^ ^ J=s fl H 3 S ^ W 1^ 54 Annals of Carnegie Museum vol. 35 External measurements Total length (Fig. 6; Table 3) is the most reliable of the external measurements of Blarina brevicauda ; it is of large enough magnitude so that it is not as readily influenced by bias as are the measurements of hind foot, length of tail and length of ear. The graph, Fig. 6, shows a general decrease in total length from north to south and from west to east. It has been demonstrated (p. 50, 52) that the mean of the sample from area A is low because of bias and it is possible that the same may be true of areas F and M. The two dimensions, length of tail and length of hind foot, did not follow the trend shown by other measurements, possibly because of the bias introduced by different collectors. The sample statistics are shown in Fig. 7-8, and Tables 4-5. A valid comparison of weights, particularly in this case, can not be made because of variation introduced by such factors as reproductive status, time of year collected, time of day collected, time elapsed since death, sex and age. Within Pennsylvania, no significant correlation can be obtained between weight and geography. Any geographic variation present is obscured by the operation of extrinsic factors, which can not be eliminated without a prohibitive amount of labor on the part of the collector. The weight of the female Blarina studied can, however, be correlated with the time of the year the samples were collected (Table 6, Fig. 9). The weights of the males, although fluctuating greatly, appeared to follow no consistent pattern. Collections were compared which were made from April through November. The weights of the females dropped from a peak in April, May and June, at which time female weight equaled or exceeded male weight, presumably in response to reproduction, to a low from August through November, when female weight averaged less than that of the males. There is one exception, area C (August-October) where the females slightly exceeded the males in weight. Evidently the influence of the reproductive cycles has a measurable effect upon the weight of female Blarina , but produces no effect, or one that is slight enough to be overshadowed by other effects, in the males. In the light of the great influence exerted by collecting bias, external measurements seemingly have slight value in a study of geographic varia- tion in Blarina , unless the entire series studied is measured by one individual, using one technique and the same instruments throughout, or unless allowance can be made for this source of error. Variations due to differing techniques of mensuration can, and in some cases do, overshadow any true variation present. Cranial measurements The cranial measurements of the specimens examined appear to fall into a definite pattern. An east-west cline is clearly demonstrable both statistically and subjectively (Tables 7-15; Fig. 10-14). Blarina crania from northwestern Pennsylvania were consistently larger in five of the nine cranial measurements considered. These measurements exhibited some degree of local, apparently random, variation, but when viewed in their Guilday: Variation in Blarina from Pennsylvania 55 entirety they decrease in size towards the east and southeast. The crania diminished in total length of skull, cranial breadth, maxillary breadth, length of tooth row, and rostral breadth. The length of the upper molar series did not follow the general trend but appeared to vary at random throughout the area. The length of the unicuspid series, however, diminished steadily toward the southeast. Width of foramen magnum, interorbital breadth and cranial height showed no apparent clinal variation. In addi- TABLE S. VARIATION IN TOTAL LENGTH OF BLARINA BREVICAUDA FROM VARIOUS LOCALITIES IN PENNSYLVANIA. (See Fig. 6) Sta- tion N M a V R A 75 116.66+0.49 4.30 3.68 108-125 B 18 123.05+0.72 3.06 2.48 116-129 C 18 121.10+0.89 3.80 3.13 115-127 D 29 119.72+0.95 5.13 4.28 106-131 E 56 120.50+0.73 5.44 4.51 106-132 F 22 114.40+0.86 4.05 3.54 108-123 G 25 118.64+1.12 5.60 4.71 107-129 H 23 119.91+1.11 5.35 4.46 110-130 I 38 117.24+0.84 5.20 4.43 106-132 1 19 118.47+1.08 4.73 3.99 109-127 K 28 121.71+0.92 4.90 4.02 115-131 Sta- tion . N M a V R L 48 119.29+0.76 5.25 4.40 110-130 M 22 111.81+1.24 5.83 5.21 102-121 N 15 119.66+0.50 1.95 1.63 110-132 O 13 117.30+1.00 3.61 3.08 112-123 P 18 118.50+1.36 5.76 4.86 110-130 Q 26 118.76+0.75 3.87 3.85 112-126 R 22 117.04+1.17 5.47 4.67 104-126 S 31 119.80+1.05 5.87 4.89 107-131 T 29 117.72+0.92 4.94 5.19 110-127 U 54 115.88+0.84 6.15 5.31 100-130 V 39 115.56+0.97 6.09 5.27 103-131 56 Annals of Carnegie Museum vol. 55 tion, the crania in the eastern portion of the state appeared more delicate in general appearance, less angular and rugged. The amount of pigment on the teeth decreased in intensity and the percentage of occurrence of the prominent hook-shaped maxillary processes became progressively less towards the east. Pennsylvania, according to the most recent taxonomic work on Blarina (Bole and Moulthrop, 1942), falls within the range of Blarina brevicauda kirtlandi Bole and Moulthrop, with intergradation in the northeastern portion of the state approaching Blarina brevicauda talpoides (Gapper). The latter is characterized as being a larger race of Blarina and if this Fig. 7. Variation in length of tail of Blarina brevicauda Fig. 8. Variation in hind foot of Blarina brevicauda Guilday: Variation in Blarina from Pennsylvania 57 is true, any intergradation should be towards increasing size in the north- eastern portion of the state. This is not the case. The animals become smaller in the east, reaching the minimum size in southeastern Pennsyl- vania. This would seem to cast some doubt on the present taxonomic status of Blarina, so far as Pennsylvania is concerned. Possibly this condi- tion occurs farther north and east, beyond the limits of the state (my data furnish no evidence on this point), but not within the state itself. From an inspection of the graphs for the various measurements (Fig. 10-14), non-clinal variation may be seen superimposed upon the clinal pattern. Examination of a few scattered localities may, therefore, give a misleading impression of the true picture. If these random variations occur in such a ubiquitous form as Blarina which exhibits no pronounced degree of ecological or geographical isolation within the area studied, they may become a problem in the taxonomic interpretation of other species which respond more readily to isolating mechanisms. This demon- strates the necessity for adequate sampling in studies of geographical variation. For example, a sample of 34 shrews from Sandy Creek, Venango County, Pa., locality F on the graphs, appears to be significantly different TABLE 4. VARIATION IN LENGTH OF TAIL OF BLARINA BREVICAUDA FROM VARIOUS LOCALITIES IN PENNSYLVANIA. (See Fig. 7) Sta- Sta- tion N M 30 - ® o x ° A O O' Id % ® <] o X <1 CO CO 20 X ^ ® O o °A CO o ® o ® A O 3 o 10 9 _ ® o A A *° Ax® OX © A 2- S1NGLES-* i i I 10 20 28 FEB. J J L 12 22 31 MARCH I I ! — ! I 21 30 APRIL -La- 11 21 3! MAY _L_ 1 10 20 JUNE ZEROS-* A Ax Ax A AA ® ® A Fig. 2. Scatter-plot. Observed numbers of individuals (ducks only) seen on spring visits in the years 1949-1952. The vertical scale is logarithmic AVERAGE NUMBER OF INDIVIDUALS (DUCKS) OBSERVED PER VISIT 1957 Preston: Waterfowl of western Pennsylvania Skyways 85 not the arithmetical difference between two observations but the ratio be- tween the two. Although there is much scatter, it is immediately obvious that a first approximation to the facts would be a parabolic curve with its axis early in April and its vertex at a few dozen ducks. Such a curve is logical. When restored to arithmetical (instead of logarithmic) ordinates, it will be a Gaussian “normal” curve. Accordingly, it seemed worth while to fit a parabola to the logarithmic plotting, by the method of least squares. In order to reduce the amount of work, all the observations over a 10-day period (the same 10-day period in each of the four years) were averaged, as in Fig. 3, and it is to these points that the parabola, the transformed Gaussian curve, is fitted. Fig. 4 shows the same data as Fig. 3 with the loga- rithmic scale eliminated and the ordinates plotted arithmetically. The curve now takes the “normal” or Gaussian “bell-shaped” form. ( NUMBER OF DAYS AFTER JAN. 31 ) Fig. 3. Averages of 10-day intervals, for the same data as in Fig. 2, plotted on the same logarithmic scale 86 Annals of Carnegie Museum vol. 35 The tentative inference from this is that the overhead traffic in the skyways pool is heaviest on April 6 (a date close to the week-end usually chosen by the Audubon Society of Western Pennsylvania to visit Oneida), but that there is a wide spread in the dates at which a fair number of ducks may be seen. In slightly more mathematical language the standard devia- tion, cr, is 24.1 days, so that one-quarter of the ducks will have passed through 16 days ahead of April 6, and one-quarter still remain to pass through 16 days after April 6. Fig. 4. Data of Fig. 3 with logarithmic ordinates eliminated Now although this is perhaps as logical and satisfactory a “graduation” as can be made of the data relating to total individuals, the picture is not at this stage complete, or might be regarded as of less than average interest, because the scaup duck (presumably in nearly all cases the Lesser Scaup) forms such a large proportion of the total population. In Fig. 5, we give the scatter-plot for number of species observed instead of number of individuals. Here we have used arithmetical ordinates, but in Fig. 6 we have graduated the logarithms, and then converted back to arith- metical ordinates in Fig. 7. The data can be fitted reasonably well to a Gaussian normal curve, as before, and its crest is at a not very different date. Since the number of species of ducks proper observable on any one day is likely to be low, we have here taken waterfowl in a somewhat broader sense, including geese, swans, grebes, loons, coots, gulls, terns, herons, and egrets, but excluding shore birds, ospreys, eagles, and kingfishers. The king- fishers are properly excluded, being summer residents. The exclusion of the 1957 Preston: Waterfowl of western Pennsylvania Skyways 87 others is somewhat arbitrary, since only the Spotted Sandpiper and Killdeer are summer residents. The exclusion of ospreys and eagles makes little dif- ference, since there is only one species of each, never more than one in- dividual, and rarely that. The shore birds also are very few in species and individuals in the spring, because the lake is full and there are no mud-flats. There is some slight possibility that the curve is bimodal, involving a preliminary small migration in February (as the ice begins to melt, on a temporary basis, at the north or inlet end, which frequently freezes), and then the main migration starting in early March. It is only the second 17- 16 15 o > 14 or Ixl £ 13 O 12 $ £ n DC 111 £ 10 8 - o 7 UJ CL 40 6 5 - or iii co 4 2 3 z 3 2 - I - xO XX x O x© x x ®x O O O O x XX A A A OO XCX A» <® ® O g SSgO X X X Q£> A O OO ® XX QDOGD AO® X A ® j A ADX ©OX® A A A X ®©® O © ® A o * 1949 x = 1950 ® = 1951 1952 <38xx O © ® X A J aJ- J 1 t 10 20 28 12 22 31 II 21 30 II 21 31 FEB, MARCH APRIL MAY Fig. 5. Scatter-plot of observed numbers of species of waterfowl in spring, 1949-1952. Arithmetical ordinates 88 Annals of Carnegie Museum vol. 35 phase that contains scaup ducks. The earlier one contains a substantial number of species, such as the Golden-eye (the “ice -duck” of the Norwegians), Red-breasted Merganser, American Merganser, Ring-necked Duck, Redhead, Canvas-back, Mallard, Black Duck, and Gadwall, as well as a gull or two. The Baldpate, though it has been recorded more than once in February on the Laboratory grounds, has not been recorded at Oneida till March 1, which is also the first date at Oneida for scaup (two individuals), but there is no other Oneida date for scaup earlier than March 11. Since observations are relatively few early in March, because I am regu- larly out of town for some days, the bimodal aspect should not be emphasized. (FEB. 1 = 1) DAYS AFTER JAN. 31 Fig. 6. Observations of Fig. 5 reduced to 10-day averages (circles). Loga- rithmic ordinates Spring migration of the scaup. Fig. 8 shows a scatter-plot of the four- year observations on this species, believed to be mostly Lesser Scaup, though, since both species (greater and lesser) have been seen at the Laboratory grounds, it is most likely that both have been seen at Oneida. Fig. 9 reduces the observations to 10-day intervals and fits a parabola to them, while Fig. 10 removes the logarithmic feature. The vertex seems to be at about April 13, which is later than for ducks in their entirety, and agrees with the qualitative observation that the scaup does not appear in February or even early March, is the last bird to linger into late May, and that big flights go through late in the season. 1957 Preston: Waterfowl of western Pennsylvania Skyways 89 Todd (1940, p. 103) however says, “The Lesser Scaup Duck moves north very early in the season,” (italics mine). Todd's work is so painstaking and thorough that any phenomena that seem to disagree with his findings should be examined with extreme care to see what the meaning of the discrepancy may be. There is no question here that his findings are right, in the sense that at least some scaup do move north very early. Todd’s earliest date is February 26, but at Toledo, Ohio, which is farther north than Oneida, it is sometimes much earlier even than this. Thus, on February 7, 1953, in company of Mr. McCormick, I found more than a thousand scaup on Lake Erie near Toledo. They were almost all males. We estimated that males outnumbered females at least twenty to one. I reported the matter to Harold Mayfield, who said that the birds must be spring migrants, since nothing like that number of scaup had been present during the winter. > (FEB. 1= I) DAYS AFTER JAN. 31 Fig. 7. Observations and curve of Fig. 6 with logarithmic scale eliminated The situation that develops then is that at Toledo the scaup is a very early spring migrant, whereas at Oneida it is one of the latest. None the less, at both places there is a remarkable predominance of males in the early movements, and perhaps later. Again, on February 3, 1953, in the company of Dr. E. S. Thomas, of Ohio State Museum, and Dr. Milton Trautman, I found at Buckeye Lake, Ohio, a few scaup, females as well as males. 90 Annals of Carnegie Museum vol. 35 Concerning this species Trautman (1940, p. 194-195) says, “The first transients (at Buckeye Lake) arrived by March 5,” but since, as above men- tioned, he was present on February 3, 1953, he agrees that sometimes the first transients are now more than a month earlier. Fig. 8. Scatter-plot. Observed numbers of scaup in spring, 1949-1952. Logarithmic ordinates 1957 Preston: Waterfowl of western Pennsylvania Skyways 91 All these observations confirm and even emphasize Todd’s comment that the scaup is very early, in the sense that the first transients of the species are very early, earlier even than he and Trautman have recorded perhaps; but so far as the average scaup duck is concerned, the Oneida evidence is conclusive that there, at any rate, the peak migration is later than that of most of the other ducks. ( FEB. 1 = 1) DAYS AFTER JAN. 31 Fig. 9. Observations of Fig. 8 reduced to 10-day intervals and fitted by a parabola The interpretation, in mathematical terms, of our joint findings would seem to be this: the scaup is by far the commonest of all our ducks, as Table 1 shows. Therefore, although the peak of its migration is late, and the standard deviation of its coming does not appear to be great (18.4 days), its very numbers give us a better chance to see something of the ends of its Gaussian distribution, and in particular to see some scaup very early. 92 Annals of Carnegie Museum vol. 35 Our own findings, that the Lesser Scaup is on the average a late, not an early, spring migrant agree with those of Kortright (1943, p. 259), “In the spring their return north is . . . tardy,” and again, “These Scaups are late breeders, next to the White Winged Scoter probably the latest.” { FEB. 1*1) DAYS AFTER JAN. 31 Fig. 10. Data and curve of Fig. 9 with logarithmic scale eliminated The fall migration. Ducks. As already mentioned, Oneida attracts fewer ducks and waterfowl in the fall, and more shore birds. The greater attraction for shore birds is obviously connected with the drop in the water-level, so that there is less acreage of water, and many acres of exposed mud-flat. However, there are probably several other factors as far as the ducks are concerned. Lake Erie, Pymatuning, Conneaut Lake, and Edinboro Lake lie from 60 to 100 miles north. Weary birds from the north would probably rest there, and when taking off again, fresh and revived, would not be attracted to Oneida. Again, in good weather, and the fall is predominantly good weather, the winds are in the west or northwest, and the birds are not likely to run into a “front” that forces them down. Thus the “net” is less efficient in the fall. Obviously the ducks are more numerous, somewhere, in the fall than in the spring, for the spring migration consists of the same birds that came south in the fall, minus those that perished over the winter. 1957 Preston: Waterfowl of western Pennsylvania Skyways 93 It is possible that the fowl use a somewhat different flyway in fall from that which they use in spring, and the fall one may not pass over Oneida; but this is conjecture and seems much less likely than that the dam is less attrac- tive and efficient in the fall. Years ago I thought that the relative scarcity of ducks on Oneida in the fall, as compared with the spring, might be because there is a hunting season in the fall. Later I came to think that it might be due to the normally much shrunken condition of the autumn lake. It does not seem likely that the latter hypothesis can be right, because F. D. Walker, of Edinboro, Pa., re- ports that on Edinboro Lake, some 70 or 80 miles north of Oneida, the ducks in spring are three or four times as numerous as in fall, though the lake is full of water at both seasons. It is a lake comparable in length with Oneida, but considerably wider. Nor does it appear likely that hunting has much to do with it, because the hunting season does not open till the crest of the migration has passed. The birds can hardly be supposed to be avoid- ing the hunting season. We seem therefore to be left with some such hypothesis as the one that assumes that these lakes are not so favorably located for the fall migration as for the spring one, possibly due to their proximity to Lake Erie, as pre- viously mentioned, or else that, on the spring migration there is some bar- rier, climatic or constitutional, to the birds’ pressing on to the north, while there is no such barrier to their southward movement in the autumn. Fig. 11 shows the scatter-plot. Fig. 12 graduates it to a parabola, and Fig. 13 eliminates the logarithmic ordinates. The average number of birds seen per visit was approximately 8, com- pared with 28 in spring. Once more, the figure refers to ducks only, and excludes swans, geese, and other water-birds. Sex ratio of the spring migrants. In my notes for February 20, 1951, is a comment to the effect that Mallards seem to come north in mated pairs, but most of the other species are predominantly unmated, and that the great majority of the birds are males. I have found occasional flocks of females, but the general picture seems to be that males are far more numerous. Whether this means that males really do outnumber females by a huge margin, or whether it means that males use Oneida in spring much more than females do, is not clear from these observations alone. At one time I assumed that the males went north to establish territories, and, running into adverse climatic conditions by starting too early, were forced to spend a little time at Oneida. The females, presumably following later, would not meet with the same obstacles and would pass right over. Possibly a count of sexes at various times on the wintering grounds in Florida or elsewhere might throw some light on the subject. Davis (1952), quoting Hochbaum (1944), says that ducks form mated pairs in the south before leaving their wintering grounds. This seems to be true of Mallards and Black Duck, but the diving ducks at Oneida in spring do not give that impression. Some diving ducks do appear in pairs, it is true, but the big flocks seem to have an overwhelming preponderance of males. The sex composition of the flocks was not always recorded, but very often it was, and a few examples are given below: 94 Annals of Carnegie Museum vol. 35 1949. Spring. Red-breasted Merganser . March 14; 5, all males. April 10; 7, 6 males and 1 female. April 27; 13, all females. April 28; 9, all females. 1949. Spring. Bufflehead. March 27; 6, all males. April 5; 12, 6 males and 6 females. 1950. Spring. Redhead. March 1; 44, 34 males and 10 females. March 12; 4, 3 males and 1 female. March 14; 10, “mostly males.” March 15; 7, 6 ZEROS X AA A 0 9 9 A A AAA 9 9 0 Fig. 11. Scatter-plot. Observed numbers of individuals (ducks only) seen on autumn visits in the years 1949-1952. Logarithmic ordinates 1957 Preston: Waterfowl of western Pennsylvania Skyways 95 males and 1 female. March 16; 8, 6 males and 2 females. March 19; 4, all males. 1950. Spring. Scaup. March 12; 7, 6 males and 1 female. March 14; 5, all males. March 15; 10, all males. March 16; 7, all males. March 19; 10, all males. March 24; 11, 10 males and 1 female. March 29; 12, 8 males and 4 females. April 11; 33, “mixed sexes.” April 13; “18 scaup and Ring-necked Ducks, mostly males”. 1951. Spring. Golden-eye. Males and females often appeared sepa- rately; when together, the males outnumbered the females. The species was present intermittently from February 16 to April 13; males from February 16 Fig. 12. Observations of Fig. 8 reduced to 10-day intervals and graduated by a parabola 96 Annals of Carnegie Museum vol. 35 to March 4, females from February 17 to April 13. Maximum number of females on any one day, 2; males, 6. Total observed males 37; females, 12. 1952. Spring. Scaup. March 13; 46 Ring-necked and scaup duck, 40 males and 6 females. March 21; scaup, 17, 16 males and 1 female. March 25; 7, 3 males and 4 females. March 29; 14, 7 males and 7 females. March 30; 14, “mostly males.” March 31; 20, 18 males and 2 females. April 6; 34, “mixed sexes.” April 13; 62, “mostly males, but some females”. 1952. Spring. Bufflehead. March 30; 23, 18 males and 5 females. 1950. All species of genus Aythya. March 1 to April 13. Approximately 175 males and 50 females. One-third of the females did not appear till April, but only about one-fifth of the males appeared in that month. There is a definite statistical pattern here, but its biological meaning is obscure. ( AUG. 6=1) DAYS Fig. 13. Data and curve of Fig. 8 with logarithmic scale eliminated Trautman (1940, p. 194-195) has noticed the early preponderance of male scaup at Buckeye Lake, Ohio. He says, “The March flocks contained more well-marked males than females or males in changing plumage; early April flocks were rather evenly divided between these classes; and in late April and thereafter the flocks contained many more females and immature males than well-marked males”. These statements are undoubtedly true, but they leave the impression that, averaged over the whole spring migration, one sees females about as often as males. This is not the case. Both Dr. Trautman and Dr. E. S. Thomas (1953), who also is very familiar with Buckeye Lake and was frequently with Dr. Trautman there, agree that, averaged over the season, observations of males outnumber those of females at least four to one; and further, they are inclined to put upon it the same interpretation as mine— that the males, coming early, dawdle by the way and spend four times as long in going through as the females do, when the latter come later and keep going. Todd (1904, p. 520, and 1940, p. 103) has no very definite observations on this sex ratio question, nor does Sutton (1928, p. 74). The phenomenon is none the less a very striking one. Shore birds. The fall of 1949 was a good time for shore birds at Oneida, and led me to suppose that many species might be seen by careful watching 1957 Preston: Waterfowl of western Pennsylvania Skyways 97 in subsequent years. In 1950, however, the summer and fall were relatively wet, and the dam was not drawn down enough to attract shore birds. In 1951 it was low, and in 1952 very low, but the count of shore birds, though higher than in 1950 (when there were none) did not reach the 1949 level by any means. Shore birds are not so easily counted as ducks, and diving ducks, when diving, not so easily as surface-feeding ducks or sleeping ducks. So long as the numbers are not large, however, or the birds too active, two or more observers can always agree about ducks. A good test was provided by the swans of April 1, 1952. Mr. McCormick counted them and I counted them. His count was 131, and mine, 133, on a single trial, and considering that the birds were moving about somewhat actively, this was satisfactory, and we thought we could compromise amicably on 132. But the same number of shore birds, unless sleeping, would be very hard to estimate with a probable error of less than 10 per cent. Killdeers are the birds that appear in greatest numbers on the mud-flats, and they are continually flying about, so that it is impossible to keep accurate track of them. In consequence, in the present paper, no quantitative work is reported on shore birds. Species-composition of the spring migrants , excluding gulls , terns , and shore birds. The faunal assemblage as seen at Oneida is tabulated below. It may not agree exactly with some of the other tabulations, since I fear I have not been absolutely consistent throughout this paper in admitting, or excluding, observations by Howard H. Elliott, and may have made other minor adjustments in one section or other of the paper. There is no reason to suspect, however, that such discrepancies affect the validity of any con- clusions that are here drawn. In the present section, only birds seen at Oneida are recorded, and Boydstown, Thorn, and the Laboratory grounds are rigorously excluded. Some of the observations are by Mr. Elliott, but the great majority are my own. The report covers the bird-days for each species as accumulated over the four-year period, on observations of the spring migration only. TABLE 1. OBSERVATIONS OF SPRING MIGRANTS Common Loon ... 109 Green-winged Teal 3 Horned Grebe ... 68 Blue-winged Teal 100 Pied-billed Grebe ... Ill Shoveller 2 Double-crested Cormorant ... 1 Wood Duck 12 Great Blue Heron ... 15 Redhead 195 American Egret ... 4 Canvas-back 46 American Bittern ... 2 Ring-necked Duck 170 Least Bittern 1 Scaup 1253 Coot ... 45 American Golden-eye 95 Whistling Swan ... 167 Buffle-head 103 Canada Goose ... 174 Old-squaw 11 Mallard ... 47 White-winged Scoter 3 Black Duck ... 240 Ruddy Duck 97 Gadwall ... 24 Hooded Merganser 78 Baldpate ... 403 American Merganser 18 Pintail ... 131 Red-breasted Merganser 178 98 Annals of Carnegie Museum vol. 35 A few interesting points emerge from this table. Although the table accounts for only 3209 ducks (excluding geese and swans), the proportion of scaup, 1253, is still just about the same percentage as in previous tables, viz., 39.1 per cent. Black ducks outnumber Mallards five to one. Since the Black Duck is primarily a bird of the eastern seaboard, and the Mallard of the interior, Oneida seems here to be definitely in the Atlantic flyway where these two species are concerned. From observations made in the company of Dr. E. S. Thomas near Columbus, Ohio, I get the impression that there, only 200 miles west of us. Black Ducks and Mallards are about equal in num- bers, and Dr. Thomas believes that the immigration of the Black Ducks is a recent phenomenon. The Red-breasted Merganser is ten times as common as the American Merganser (Goosander), and twice as common as the Hooded Merganser. The Ring-necked Duck and the Redhead are about equally common (170 and 195 respectively), and the Ruddy Duck is about as common as the Buffle-head (97 and 103), or as the Golden-eye (95). Although the Black Duck is so much commoner than the Mallard, it is not the commonest of the pond ducks. That distinction belongs to the Baldpate (403 vs. 240), but this is due to the Baldpate often appearing in very large flocks, occasionally of 60 or 80 individuals. The Black Duck is seen more often, but the largest flocks observed were of 16 and 20 individuals. Baldpates were seen (in spring) on only 34 occasions, while Black Ducks were observed 57 times. With those species that move in large flocks, a single observation, made or missed, can make a vast difference in the count. Thus of the 167 swan-days reported, 157 are due to the single large flock of 132 birds forced down by a line-squall, as previously mentioned. Half of the goose- days are due to a single flock of 85 birds observed by Howard Elliott one evening. Superficially, there is a great difference from year to year, even with those species where the total count is not due primarily to a few large flocks. Thus the annual spring count for scaup goes: 140, 466, 353, 294; but for Ring-necked Ducks the progression is quite different, viz. 81, 52, 16, 21. The count for Black Duck was 77, 99, 57, 7; and for Mallards, 12, 19, 16, 0. These differences do not seem to be due to the relative frequency of my visits to Oneida, and it is not clear that they are due to timing, though both may play some part. Black Duck seen per visit (counting only those visits on which at least one was seen) come out as follows: 1949, 4.5; 1950, 4.7; 1951, 3.8; 1952, 1.7. But it is possible that the differences are not significant, and it seems improbable, in view of the tendency of this species to move at times in pairs and at others in flocks, that any simple statistical treatment would help. The Common Loon. It is likely that, even when a species is not ob- served on more than a few dozen occasions, or in numbers exceeding 1957 Preston: Waterfowl of western Pennsylvania Skyways 99 a few score, a detailed count on each occasion may permit the testing of various hypotheses or current notions on how the bird behaves. This has been done in some detail in a separate paper (Preston, 1956)*, and it is therefore unnecessary to do more than summarize it here. The Common Loon is an extremely late migrant in spring, lagging far behind the ducks as a whole and even behind the Lesser Scaup. The peak of its migration is at April 22. Further, this species migrates much more nearly simultaneously than ducks as a whole, or even than the scaup. The standard deviation of the timing is 11.6 days for the Loon, 18.4 for the scaup. This suggests that the Loon does not start north till its breeding grounds are ready for it; and when it starts, it goes right through. Thus the Loon is probably commoner in the skyways than its numbers, compared with the scaup, would suggest. With a cruising speed, unhurried, in excess of sixty miles an hour (Bent, 1946; Preston, 1951), it does not need to dally at Oneida. We see few of them unless they are forced down by fog or rain. Normally the Loons we see at Oneida are not organic flocks, but random assemblages. At the peak of the season the most likely number to be seen is three, and they are usually well separated and paying no attention to each other. There is no evidence in the field that any of them are mated pairs, and a mathematical analysis of the subject suggests that they are moving as individuals, not as pairs. Table 2 gives a conspectus of the migrant water-birds seen, in spring and fall (S and F) for the four years 1949 to 1952 inclusive. It is quantita- tive only as to species. Some species may be represented by only a single individual, as with the Northern Phalarope, and others by more than a hundred, as with the Whistling Swan on April 1, 1949. A few comments on Table 2 may be useful. First of all, some species are conspicuous by their absence, and others have not shown up to advan- tage in this four-year period. Thus, I once saw a large flock of Snow Geese on Oneida under the thick mists of early morning, and Mr. Norris has reported a large flock of Snow Geese flying north over the Laboratory grounds. On November 16, 1948, J. M. McCormick found two immature Blue Geese at Boydstown Dam, immediately above Oneida, and it is reasonable to suppose this species occasionally visits Oneida. Only one, slightly uncertain, record is given for Oneida itself, the bird disappearing before I could be absolutely sure of it. Again, only one record is given for the Caspian Tern, in the spring of 1950, but another individual (or the same one) spent some time there in the fall of 1948 (September). The Dunlin, or Red-backed Sandpiper, was present in the fall of 1948, too. The Shoveller is recorded on Oneida only in 1950 (spring); but it oc- curred on other, smaller, local ponds once or twice in this four-year period, and I saw it in greater numbers on Oneida a few years earlier. The Canada Goose probably occurs every spring and most falls, but it is somewhat easily scared out and probably remains only from dusk to dawn. On one •Charles S. Pearce, Manager of the American Ceramic Society, of Columbus, Ohio, says that paper should be entitled “The marital status of migrating loons”. 100 Annals of Carnegie Museum TABLE 2. MIGRANT WATER BIRDS 1949 1950 Species S F S F Common Loon x x x x Red-throated Loon x Horned Grebe . . x x x Pied-billed Grebe x x x x Double-crested Cormorant x Great Blue Heron x x x American Egret Little Blue Heron x American Bittern x Whistling Swan x x Canada Goose x Blue Goose x Mallard x x x x Black Duck x x x x Gad wall x Baldpate x x x Pintail x x x Green-winged Teal x Blue-winged Teal x x x Shoveller x Wood Duck x Redhead x x Ring-necked Duck x x x Canvas-back x Scaup x x x American Golden-eye x x Buffle-head x x x x Old-squaw x x White-winged Scoter Ruddy Duck x x x x Hooded Merganser x x x American Merganser (Goosander) x Red-breasted Merganser x x x American Coot x x x x Semipalmated Plover x Killdeer x x x Wilson’s Snipe Solitary Sandpiper Greater Yellow-legs x x x Lesser Yellow-legs x x Pectoral Sandpiper x White-rumped Sandpiper x Baird's Sandpiper x Least Sandpiper x Red-backed Sandpiper (Dunlin) x Stilt Sandpiper x Semipalmated Sandpiper x Northern Phalarope x Herring Gull x x Ring-billed Gull x Bonaparte’s Gull x x Common Tern x Caspian Tern x Black Tern x Forster’s Tern x 1951 x x X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X vol. 35 1952 X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X * * H K X X 05 1957 Preston: Waterfowl of western Pennsylvania Skyways 101 day in the quadrennium, Howard Elliott saw a large flock come in at dusk half an hour after I had left Oneida. In 1952 a single individual came in to the Laboratory grounds at 4:30 P.M. and stayed into the night but was gone early next morning. The Black-crowned Night Heron presumably occurs regularly at Oneida, but is abroad only at night, when I am not there. I have seen the species at Thorn Dam at dusk, and at the Laboratory grounds for weeks on end. Forster’s Tern is a new species for western Pennsylvania, and Mr. Todd, properly, rejects the observation. I did not at the time realize that it had never been seen here before, or I could perfectly well have collected it, but since Mr. McCormick and myself are both completely satisfied, I can not very well reject my own records. The bird may be seen in some numbers at Toledo, Ohio. The Black Tern is probably more frequent than the records indicate. It has several times been seen on the Laboratory grounds and at even smaller ponds. Bonaparte’s Gull seems to be regular in spring, but absent in the fall. The other gulls and terns are seen at both seasons. The Canvas-back and Old-squaw also seem to be spring migrants, while the other ducks, geese, and swans are likely to show up in the fall also, though not in such numbers or with equal certainty as in the spring. I have expected the Gray (Black-bellied) Plover and the Golden Plover to show up in the fall on the mud-flats, some of which are wet and some fairly dry, but I have never seen either species, either in the quadrennium or earlier. They occur in some numbers near Sandusky, and near Columbus, and in great numbers on the New Jersey marshes, so presumably one of these days a few strays will reach Oneida. At present I am inclined to suspect that the skyway over Butler County does not normally contain this species at all, but there may be other explanations. Near Toledo and Sandusky the birds are remarkably local, though very plentiful where present at all.* Of the other birds that might be expected to be attracted to Oneida, some mention must be made. The Bald Eagle is occasionally there but only as a bird of passage and very briefly. I have seen it attacked by crows. The Osprey is much more regular. In fact, in April it is present all the way from Oneida northeast to New York state. Apparently it catches fish in the riffles of the streams, from the Allegheny River at the junction of the Clarion over to the Susquehanna watershed. It regularly appears at the Laboratory grounds, and catches fish in the Carrie Dam. Occasionally it is present as late as June. The Belted Kingfisher is present about nine months of the year, and nests (or has nested) below the Oneida spillway. The Little Green Heron is present on all the streams, probably, of the district, and regularly nests at the Laboratory grounds. It is, therefore, omitted from the tabulation. The Spotted Sandpiper nests at Oneida, as well as at the Frith (the inclosed grounds of the Preston Laboratories, three miles due west of the center of Butler, Pa.), but at the former place it appears to me less common than it was years ago. This may be because of the great increase in the •A single individual of the Black-bellied species finally put in an appearance on September 29, 1953. 102 Annals of Carnegie Museum vol. 35 number of fishermen that crowd the banks. The Solitary Sandpiper is more common at the Laboratory than at Oneida, perhaps because it is a woodland bird. The American Merganser is occasional at Oneida, but probably regular on our swift-flowing streams, such as the Slippery Rock Creek in the Main Gorge, and the lower Connoquenessing above its junction with the Slippery Rock. Only one American Egret has been seen at Oneida on migration proper (spring of 1951). Five individuals were present at Oneida from August 8 to September 19, 1948; and, many years earlier, about 1928, I saw about a dozen of them there. An adult bald eagle that flew over the Laboratory grounds at a great height on May 13, 1951, headed due north, was perhaps a Florida eagle headed for the Great Lakes or the Canadian north woods. I myself have never seen the Florida Gallinule at Oneida, but a bird of the year was caught there by one of my assistants (Sept. 14, 1951). The rarity of the bird in Butler County is a great surprise to me, since in England, under the name of moorhen or water-hen, it is the most widely met of all waterfowl. Of the passerine birds there is little to say, since few of these are water- birds in any real sense. Oneida is a good place to see the early swallows, and on the cold spring mornings the Carrie Dam in the Frith is a good place, too. The water supplies the few insects that are abroad when the ground is chilled to the freezing point but the water is open. The American Pipit is a tundra bird. In the fall I have seen it in hundreds among the dead tree stumps of the mud-flats of Mosquito Creek Dam, just over the border in Ohio. The only place I myself have seen it in Butler County is at Oneida Dam, on two occasions, once in some numbers. But Howard Elliott and Earl Schriver have reported flocks of the species in other parts of the County, so neither water nor mud-flats are essential. Crows are rather plentiful around Oneida, and in the fall vast flocks of Starlings, with some Red-wings and perhaps other species, roost in the woods adjoining the lake, but this has no obvious connection with the presence of the lake itself: it is not the only Starling roost of the County, nor a principal Crow roost. The Rusty Blackbird occasionally appears at the Laboratory, but I have not yet recorded it from Oneida, where it might be expected. Rarities , unusual dates, etc. Birds not often seen in western Pennsyl- vania, at any rate not often seen south of the lakes area in the extreme northwest corner, or reported by Todd (1940) to be infrequently observed there, are as follows: The Red- throated Loon (November 15-17, 1949) has been reported very seldom. One of the earlier records, however, was from Oneida. The Double-crested Cormorant (April 13, 1950) is quite uncommon, though one made itself conspicuous along the Allegheny River a few years ago. The Whistling Swan is common enough, and 1 mention it here because its flyway appears to be narrow. The bird is common near Youngstown and on Lake Erie near Toledo, but unusual near Columbus, Ohio. It would seem that Columbus lies southwest of the flyway. The Gadwall is fairly regular in its appearance. Todd seems to have thought it rather uncommon short of the Lakes. 1957 Preston: Waterfowl of western Pennsylvania Skyways 103 Todd says the Long-tailed Duck, or Old-squaw, used to be common on Lake Erie but has been all but exterminated in the interests of the fisheries. None the less it has visited Oneida every spring in the four years under review, though only in small numbers. The White-winged Scoter is reputed to be regular at Youngstown, but has been reported only twice at Oneida, both times in the spring of 1951. The Northern Phalarope has appeared twice, and the 1952 individual stayed for about a week. The bird is regularly seen near Toledo, but not in any numbers; it is rare at Buckeye Lake, Ohio. It may be less rare overhead in Butler County than at present appears. The Caspian Tern and Forster’s Tern have already been commented on. The Caspian Tern’s date was very early (April 11, 1950). On the previous day, driving home from Toledo, I had seen a couple of these birds at a small pond near Fitchville, Ohio. A very strong west wind was blowing, and I thought it possible some migrants might be blown into Pennsylvania. I suggested to Mr. McCormick that we go to Oneida and find a Caspian Tern. We went, and there was the bird. The Little Blue Heron was, of course, an immature in white plumage. The same species, in the same plumage, has been seen on the Laboratory grounds, but only once. The White-rumped Sandpiper is represented by a single individual at Oneida, but again, a single individual was also seen in an earlier year, at the Laboratory. The Greater Yellow-legs stays somewhat later than Todd reports. In 1949, one individual was present on November 20, an all-time record accord- ing to Todd, but 20 were present on November 10, and eight on No- vember 15. Three Dunlins (Red-backed Sandpipers) were present on November 20, 1949, again an all-time record according to Todd, but one was present on November 24, Thanksgiving Day. The Dunlin is a very late migrant at Toledo, and I think Todd underestimated its tendency to be late on the fall migration. It is the last of the Sandpipers to go through, and it out- stays even the Greater Yellow-legs. Three Killdeers remained till December 4 in 1949, which is quite late with us. In 1950, a male Hooded Merganser appeared on March 15, which is a comparatively early record. On April 20, 1950, Mr. McCormick and myself found at Oneida two Greater Yellow-legs with one Lesser Yellow-legs with them, so there was no question of its size. It is a very early record for the latter species. On February 16, 1951, I observed four male Golden-eyes and one male Canvas-back. The Golden-eyes were early, and the Canvas-back three weeks earlier than anything reported by Todd. The first Ring-necked Duck, a female, appeared on February 23, 1951, which is very early, but a male had been at the Carrie Dam in the Frith since February 19. Eleven Redheads appeared on February 28, 1951, a day or so earlier than Todd’s records, but agreeing closely with our own date of March 1, 1950. I 104 Annals of Carnegie Museum vol. 35 The American Egret of April 15, 1951, was a somewhat unusual observa- tion, spring visitors being rare in Western Pennsylvania. Pectoral Sandpipers remained at Oneida in 1952 until October 26. From the detailed records, it is possible to tabulate the first and last (observed) appearance of each species on both spring and fall migrations within the period of four years, but, though such information is not too plentiful, especially as regards last appearances, it is doubtful whether it has much utility. At the best, it represents only the tail-ends of more or less Gaussian distributions, which theoretically extend indefinitely to left and right. There is a good deal of evidence in Todd (1940) that many species of ducks will stay on Lake Erie all winter if the winter is exceptionally mild, and a good deal of evidence that many unexpected species will stay and breed at Pymatuning if they find conditions to their liking. Both these facts emphasize the Gaussian character of the time “limits” of the spring and fall migrations. The birds are not set on reaching a particular geo- graphical region, but only on reaching a satisfactory environment. If we could maintain a suitable environment all year round, the birds might cease to migrate at all. This is not altogether surmise. In Britain, with its equable climate, the birds tend to be much less migratory than the conspecific forms of continental Europe. The recent tendency of the European White Stork to nest in South Africa, hitherto merely its wintering grounds, may be an example of the same sort; and the colonization by the Gadwall of the Outer Banks of North Carolina (Pea Island), a thousand miles from its normal prairie summering grounds, indicates the same thing. Migration may be a deeply ingrained instinct, but it seems to be one that is readily dropped as soon as it ceases to be a painful necessity. UTILITY OF THESE OBSERVATIONS It is highly probable that if Oneida had been watched continuously over this four-year period, or if it had even been visited every day, a number of additional species would have been seen, and many more significant observa- tions made. As it is, it is probable that no small body of water in our region has been watched equally closely or quantitatively. None the less the period is short and a decade might have been more useful. However, it is probable that the picture is changing. There are vastly more small farm ponds on the Foreland than there were 10 or 20 years ago, and these attract success- fully many species, including birds as large as geese, and ducks as relatively uncommon at Oneida as the Shoveller. The number of such ponds will undoubtedly increase, and some may prove superior to Oneida in their seclusion, for Oneida is much disturbed by fishermen and other visitors. When Oneida was new, thirty years or so ago, and less thickly surrounded by anglers throughout most of the year, it is probable that it was more visited by waterfowl and waders (shore birds), especially in the fall. Unfor- tunately, we have no quantitative records, known to me, of those days. It seems to me that, if Oneida is watched long enough, everything that has ever been seen at Pymatuning or Erie, in the line of waterfowl, will be seen also at Oneida, and perhaps some species that have not been seen farther north. If this is so, it seems clear that the proposed lake in the valley of Muddy Creek, a few miles to the west, will be a far greater attraction; 1957 Preston: Waterfowl of western Pennsylvania Skyways 105 for Oneida is a reservoir of drinking water, and is not operated with any eye on the needs of water-birds or the interests of bird lovers or ornitholo- gists. The lake is small, largely free of vegetation, quite free of islands, and, on the face of things, quite unattractive to birds. The proposed lake in Muddy Creek would be many times as large, much more complex in out- line, with much more extensive shore -lines, and would intercept a much greater width of flyway. Whereas no part of Oneida is free from frequent, almost incessant, dis- turbance from human beings, being completely and closely surrounded by roads and, for most of the year, lined with fishermen, there is a possibility that Muddy Creek can, at least in part, be set aside for wildlife. At Oneida, while with Mr. McCormick, I have seen a fine antlered deer swimming the lake, and with Mrs. Preston have seen a small herd of deer come down to drink. On Boydstown, beaver could sometimes be seen, and there were until recently several beaver dams upstream from Boydstown. Both species of mammal still occur in Muddy Creek, and there also I have seen the Bald Eagle, the Bittern, Virginia Rail, American Egret, and many other species, though at present there is no expanse of water there. The Marsh Hawk and Red-shouldered Hawk (a swamp species) both nest there, and I have never seen either at Oneida. The Snow Goose has been reported on Muddy Creek, and the Wood Duck and Mallard nest there. Howard Elliott has reported the King Rail nesting at a point not much beyond it, and it would seem that Oneida is only a very tattered net for sampling our skyways. METHOD OF MAKING COMPUTATIONS Fig. 2 shows the scatter-plot representing all observations, in four spring- times, of the ducks (excluding geese and swans) observed on the water (not overhead) at Oneida. Such a scatter-plot really contains more information than any other plot that can be devised, but it is too complex to be grasped, and therefore lacks utility. Fig. 1 1 shows a similar plot for the four autumns, and the same comments apply. It is difficult to make out what is going on. Accordingly, it is convenient, and useful, for some purposes, to reduce the number of plotted points to a modest number of average points. This has been done by taking 10-day periods, the first being February 1 to 10, inclusive; the second, February 11 to 20, and so on. February 29, 1952, is treated as non-existent, so that March 1 is in every year treated as the 29th day, and March S as the first day of the fourth 10-day period. Since observations were not made on every day, it may happen that the average time of observation during the 10 days does not come at the mid-abscissa of that period. Its true position is calculated, and the average of the number of birds seen on each visit constitutes the ordinate for that period. By this means we reduce the plotted points to fifteen or so, and the general shape of a well-fitting curve becomes apparent. (Fig. 3, Fig. 12, etc.) In Fig. 3 the circles are observed points, and the figure suggests that the cut-off in late May is sharper in practice than the computed curve would imply. However, very few visits were made in late May and in June, so the evidence is here incomplete and the computed curve is possibly fairly accurate. 106 Annals of Carnegie Museum vol. 35 Curves are sometimes fitted on a purely empirical basis. This is not a particularly useful proceeding. The curve that is to be fitted should be logical, not merely a good fit; and of two curves that fit reasonably well, the more logical should be chosen even if it is a slightly poorer fit than the other. The question of logic involves a reference to the physics or biology of the problem and can not be decided solely by statistical ideas. Thus, the fact that we see a period of maximum abundance of ducks on Oneida near April 6 and again near October 16 does not constitute all that we know about the skyway pool. We know that the birds in spring are headed one way (conventionally “north”) and in the fall are headed the opposite way (conventionally “south”). If we take this vector knowledge into account, the question we are really asking is: “How full is the skyway of birds flying north?”, or “How much northing is the bird population over- head accomplishing per day?” When expressed in this form, the fall migrants are negative birds, because they are losing “northing.” If spring and fall migrants are plotted on the same diagram, the fall migrants should be plotted below the zero line (Fig. 14). We are then led to treat the annual migration as a single event, and should logically “graduate” the curve as a sine curve with a periodicity of 12 months. Such a sine curve has some ele- ments of logic, but the diagram rather discourages it; there is too long a period in the summer, and too long a period in the winter, with zero birds. The sine curve would pass through zero twice a year, but the zeros ought to be short and sharp. The observations themselves accordingly discourage this concept, and this discouragement is fortified by our knowledge of their biology. The birds are not flying north or south to accommodate themselves to the rhythmic movement of the isotherms. They are moving north to breed and moult, a business that takes time, and so there is naturally a complete absence of birds for some time in the summer. In the winter they have to eat, and they will naturally be absent from the frost and freezing, till open water becomes dependable. On this basis, the birds’ biological requirements could be met by the whole of them moving north simultaneously on April 6, and all moving south on October 16. This would be similar to the daily “rat race” that develops in Pittsburgh and other large cities, when all the suburbanites crowd into the down-town districts about 8 A.M. and all rush out at 5 P.M. The birds also spread or “stagger” their comings and goings over a sub- stantial period. This arrangement presumably arises from the fact that the birds are not exactly synchronized in their internal biological rhythm, and they are not all flying the same routes or starting from the same points. Accordingly, a reasonable assumption would be that enough variable fac- tors are present to induce in the observations the appearance of a “normal," or Gaussian, “curve of errors.” The two migrations, spring and fall, are then to be thought of as separate events, even though one does involve the other. But since the birds do not necessarily follow the same routes coming and going (a matter we know for certain with some other species), and since their behavior, and the com- position of the flocks, is most likely quite different at the two seasons, we 1957 Preston: Waterfowl of western Pennsylvania Skyways 107 are justified in regarding the two events as independent problems, and graduating each one with a single Gaussian curve. Consider the spring migration. Let its peak occur at £ days after January 31. Let the most probable number of birds to be seen on a single visit at the peak be tj birds. Let us calculate the most probable number ( N ) of birds to be seen on any other occasion, dated x days from January 31. (x — £)a We have N ~~ rj . e (1) where C3 CO CO o CO PM 1 > > CM QJ rfM CM CM CJ o £ Dm W u Dm > o Dm CJ CM UO o co CO o . — ' O O QJ CO Z CO 05 05 05 o o CM CO c~~ o >- .CM CM CM CM 05 CM CO *4 o3 05 Ms c3 03 Pm < Sh 03 M<‘ 03 Jh o3 5-J c3 o 5^ CM * s s s a CM Dm QJ CO O CM L 03 CM ^ co CM 00 CO s z ri 121 i? z o 50 P < eo . > > O o *7- Oh QJ CO CO CO CM IT) CO o 03 1—1 CM CM CM >> Oj Ms’ 5* Dm^ •D QJ ^ Dm Ms’ Dm J-1 03 .D QJ QJ s C PM < < pM pM xh ^ > 00 > • 0.0 tJ S5 QJ O CO w CM o 05 m CM ^ CO CO o ^ ® ° z !>■ rf CM ™ > oz t - OO M 00 04 OD « , * CM ^ CM ^ It < H 50 3 < 05 05 >—• iT5 CM CM CM —i >-n •— M 03 03 U O o o ^ > Pm CM mJ CJ > o 2: mJ u DmO QJ u M.O M-M Dm Dm w u o ^ o ^ z Dm 05 o co QJ QJ QJ QJ TjH CO CO CO CO 05 CO m CM CM 05 OO o — < CM CM *”• f-J CM CM CM co Ms Dm < Ms Dm < S3 2 Apr. aF JD .H x> QJ -O Pm 50 3 < _ OO so qj • »“M u QJ Dm CO 3 O O H-! 3 O e a o U qj •D qj Mm o w (0 -Q QJ (m o T5 qj »— < • »-* -a qj S Green-winged Teal Nov. 15 Mar. 21 Oct. 28 Sept. 21? Nov. 20 May 6 Nov. 4 1957 Preston: Waterfowl of western Pennsylvania Skyways 113 © © © Pn *-< CM CM . cm CL CU tS tj CM © cu co o o o OO © CM cl < u o u c © OO © © © 04 © r-1 co CM f-H © CM © >-© CM CO CM r^ $4 © 5-1 CM »4 . CL • rS pS O, . CL CL .d •g < £ S xi v4 a &H* n . s4 PS s < -f < s4 cS < <3L> Pl QJ L-U f* £ qj Pl 2 Wh < 2 s cu ptl 2 $ > O ^ O Z « °r n. QJ QJ ' to 0/3 PL —* bD * O ; Z CO CO . > > o ° z 3 < ^CM f-H CM ’“H v4 S_< CM v4 t4 £ Ma Feb. X3 QJ fr4 Ma Feb. L CL CL < CL < >4 Pj CL . CL < a.< o. rt 2 2 Uh 2 < CJ cu 2 < < 00 > M- ^ O X Z X « CL ° Jj © -> a/ Ui f-H OO F-H oo oo f-H CM f-H CM CM ilO © > > > > CM > > o > o > o > O o 0 Z z o z Z o Z z 0 & Z Oct Z t" oo c-~ 00 © © CM CM © CO 1 © © ^ CM ’~i re 10 © © CM < f-H CM S-.CM f-H *4 .CM CM ^ CM CU «( C3 a CL • C^! nS PS >4 ctf S M »4 ps < >4 ps s Jh s4 S 1 Cl, J§ ;4 PS < S3 (4 pS y »4 F=i fX, ^ CL 2 2 2 s < < s S < < © : s* ^ > OO CM © 0O ^ s* $ d8 > Q > O o Z Z 2c £ 00 & CO 27 © CM 00 CM © CM 25 27 oo t4 ct © <4 fS >4 PS xi s4 cS Apr. 4> Pl s 3 cu XS < Sd JJ £ -2 *53 > Bonaparte's Gull Mar. 24 Mar. 29 Apr. 1 Apr. 13 Apr. 23 Apr. 20 114 Annals of Carnegie Museum vol. 35 These results may perhaps be usefully condensed to give the earliest and latest dates so far observed for migrant species in spring and fall, as follows: Common Loon March 19— May 18 October 19— November 25 Horned Grebe March 28— April 23 November 13— November 23 Pied-billed Grebe February 16— May 12 September 1— November 24 Great Blue Heron March 22— April 27 August 12— October 22 Whistling Swan March 29— April 25 October 29— November 17 Canada Goose March 19— March 27 October 13 Mallard February 12— May 1 September 10— November 11 Black Duck February 12— May 14 August 30— November 20 Gadwall February 20— May 11 September 3 Baldpate February 21— May 1 September 14— November 18 Pintail March 15— March 20 August 29— November 18 Green-winged Teal March 21— May 6 September 21— November 20 Blue-winged Teal March 25— May 12 August 12— November 19 Redhead February 9— March 1 November 18 Ring-necked Duck February 9— April 27 October 18— November 11 Scaup March 1— May 18 October 20 — November 23 American Golden-eye February 9— March 12 October 20— November 23 Buffle-head March 21— May 1 October 28— November 27 Old-squaw April 6— cApril 25 (no Fall observations) Ruddy Duck March 21— May 13 October 17— November 23 Hooded Merganser March 13— April 20 October 4— December 4 Red-breasted Merganser February 9— May 1 October 28— November 20 1957 Preston: Waterfowl of western Pennsylvania Skyways 115 Coot PSeptember 10— November 24 Greater Yellow-legs April 9— May 14 PSeptember 1— November 20 Lesser Yellow-legs April 20 Bonaparte’s Gull September 10— PSeptember 28 March 24— April 23 (no Fall observations) 116 Annals of Carnegie Museum vol. 35 REFERENCES Bagg, A. M. 1950. Barometric pressure patterns and spring bird migration. Wil- son Bulletin, v. 62, p. 5-19. Bennett, Holly Reed 1952. Fall migration of birds at Chicago. Wilson Bulletin, v. 64, p. 197-220. Bent, A. C. 1946. Life histories of North American diving birds. 237 p. Dodd Mead & Co., New York. Davis, D. E. 1952. Social behavior and reproduction. Auk, v. 69, p. 171-182. Gunn, W. W. H., and A. M. Crocker 1951. Analysis of unusual bird migration in North America during the storm of April 4-7, 1947. Auk, v. 68, p. 139-163. Hald, A. 1952. Statistical tables and formulas. 92 p. Wiley, New York. Hochbaum, H. A. 1944. The canvasback on a prairie marsh. American Wildlife Insti- tute, Washington, p. 1-201. Kortright, Francis H. 1943. The ducks, geese and swans of North America. 476 p. Ameri- can Wildlife Institute, Washington. Mayfield, H. F. 1947. Spring migration, Ohio-Michigan region. Audubon Field Notes, v. 1, p. 153-154. Preston, F. W. 1951. Flight speed of common loon. Wilson Bulletin, v. 63, p. 198. Preston, F. W. 1956. The migrant loons of western Pennsylvania. Auk, v. 73, p. 235-251. Sutton, George Miksch 1928. The birds of Pymatuning Swamp and Conneaut Lake, Craw- ford County, Pennsylvania. Annals of the Carnegie Museum, v. 18, p. 19-239. Thomas, Edward S. 1953. Personal communications. Todd, W. E. Clyde 1904. The birds of Erie and Presque Isle, Erie County, Pennsyl- , of the Carnegie Museum, v. 2, p. 481 -596-. Todd, W. E. Clyde ’ ’ 1940. Birds of western Pennsylvania. 710 p. University of Pittsburgh Press, Pittsburgh. Trautman, Milton R. 1940. The birds of Buckeye Lake, Ohio. 466 p. University of Michi- gan Press, Ann Arbor. £ O ?4 * ^ P h MAY 2 0 1959 ART. 8. SYSTEMATIC NOTES ON NORTH AMERICAN BIRDS 2. The Waterfowl (Anatidae) By Kenneth C. Parkes Associate Curator of Birds, Carnegie Museum This is the second of a series of papers on the systematics and nomencla- ture of certain North American birds. A general introduction and acknowl- edgments for the series as a whole will be found in the first paper (Parkes, 1955c). The waterfowl have long been among the most popular birds, and there has been a recent renewal of interest in their classification. The now classic paper by Delacour and Mayr (1945), in which a radically new alinement of the swans, geese and ducks was proposed, may be regarded as a major turning-point in waterfowl taxonomy. Most of the changes from the tradi- tional classification employed by Peters (1931), even though sometimes quite drastic, have been accepted by most students of this family. In the decade since the revised classification of the waterfowl appeared, many papers have been published which supplement or correct some of the state- ments made by Delacour and Mayr, or take issue with them on specific points. Differences of opinion are to be expected among taxonomists, and Delacour and Mayr would be the last to claim that theirs was the final word. In general, however, it may be said that our knowledge of the systematics of the waterfowl, when compared with other bird families of similar size, is exceptionally thorough. Special mention must be made of the work of Verheyen, who has taken sharp issue with Delacour and Mayr, and proposed his own classification of the waterfowl (summarized in Verheyen, 1955b) based primarily on com- parative osteology. A thorough critique of Verheyen’s work may be accom- plished only by a comparative anatomist. It is quite apparent, however, that Verheyen has fallen victim to the tendency described by Simpson (1945, p. 23) as “the tendency to raise the ranks of groups without need, that is, without gaining any practical advantage. One of the more evident symptoms of this tendency is the appearance of many monotypic groups in classifica- tion. If a classifier makes mostly monotypic families, genera, etc., it is a fair statement that he is giving family rank to what should be called genera, generic rank to species, etc.” That this is descriptive of Verheyen’s classifica- tion is made plain when it is pointed out that he has divided the relatively homogeneous family Anatidae into sixteen families, no less than six of which are erected to include a single species! Brief mention should also be made of the work of Yamashina (1952), who proposed a classification of the Anatidae based entirely on cytological and hybrid sterility data. Although he examined only fifty species, his classi- fication agrees in many respects with that of Delacour and Mayr. Yamashina, however, has “lumped” genera to an even greater extent than did Delacour and Mayr, and his classification represents the opposite extreme in taxonomic practice from that of Verheyen. Submitted for publication, November 15, 1957 Issued November 7, 1958 117 INSTITUTION A™* * 15,3 118 Annals of Carnegie Museum vol. 35 Since the appearance of the first paper in the present series, the long- awaited fifth edition of the American Ornithologists’ Union “Check-list of North American birds” has been published (1957). The classification of waterfowl employed by the A.O.U. is a conservative one, based chiefly on that of Peters (1931). Reference will be made beyond to some of the ques- tions on which I feel that the A.O.U. Check-list is conservative to the point of being reactionary. 1. The Swans The classification of the swans hinges on the relative importance assigned to certain osteological features. Although Delacour and Mayr (1945) omitted all reference to the striking internal differences among these superficially similar birds, these structural features are described in Delacour’s recent book (1954, p. 57, 71). Wetmore (1951) believed that two genera should be recognized; Cygnus for those species in which the trachea passes directly into the thorax without entering the sternum, and Olor for those in which the trachea loops into the sternum, the furculum being modified at the sym- physis to accommodate this loop. In addition to these major anatomical features, the two genera may be separated by certain relatively minor ex- ternal characters, such as the cuneate tail of Cygnus versus the rounded tail of Olor. The two groups also differ in behavior patterns, and it seems worth while to follow Wetmore in recognizing two genera. There are also two schools of thought regarding specific limits within the genus Olor. Delacour and Mayr (1945, p. 8) followed the suggestion of Hartert (1920, p. 1275) that the New World O. columbianus and O. buccina- tor be considered conspecific with the Old World O. bewickii and O. cygnus , respectively. In the case of the Whistling Swan (O. columbianus ), such action is justifiable. A glance at the range map published by Delacour (1945, p. 84-85) shows that the New World columbianus and the two Old World forms bewickii and jankowskii are obvious geographic representatives, dif- fering slightly in size and conspicuously in the relative amount of yellow at the base of the bill. All three are highly migratory Arctic nesters. I there- fore agree that these two Old World forms should be considered subspecies of Olor columbianus. The second case, that of O. buccinator and O. cygnus, is quite different. Here Wetmore (1951) and Delacour (1954, p. 71) differ on a question of fact rather than merely on interpretation of accepted facts. Wetmore recog- nized the subgenus Clangocycnus Oberholser for the Trumpeter Swan, O. buccinator, basing this on the fact that in this species the trachea makes “a dorsal loop as it enters sternum, protected by a bony case that projects into the anterior end of the body cavity.” Delacour (1954, p. 71) claims that this is also true of the Whooper Swan, O. cygnus, which he therefore combined with buccinator as a single species. Wetmore, on the other hand, placed cygnus in the typical subgenus Olor. I have examined sterna of the swans in question, and find that Wetmore is correct; the sternum of O. cygnus does not have the enlarged “bony case” typical of O. buccinator, but is somewhat intermediate toward that of O. columbianus. This is illustrated by Schiller (1925, pi. 65). 1958 Park.es: Systematic Notes on North American Birds 119 As mentioned above,- Olor columbianus and O. bewickii are clearly geo- graphic representatives of one another. This can not be said for O. buc- cinator and O. cygnus * The Whooper Swan is primarily an Arctic nester, although it does extend into central Asia. It is strongly migratory. The Trumpeter Swan, on the other hand, is decidedly a bird of temperate re- gions, and is relatively sedentary. The Whooper Swan resembles Bewick’s Swan in having a large area of yellow at the base of the bill, a feature of which there is no hint in the Trumpeter Swan. The total of the evidence seems to suggest that Olor cygnus is a larger derivative of the same stock which gave rise to the O. columbianus group, and is no more than generically related to O. buccinator. 1 believe, therefore, that the two North American swans should be known as Olor columbianus columbianus (Ord) and Olor buccinator (Richardson), respectively. This differs from the current treatment in the A.O.U. Check- list only in considering the Whistling Swan to be the New World representa- tive of a polytypic species. 2. The genus Branta Verheyen (1955a, p. 9) has introduced into the literature the name Eubranta , apparently as a new genus intended to include the Barnacle Goose, Branta leucopsis, and the Red-breasted Goose, B. ruficollis. He gives no diagnosis of this new genus, apparently basing it primarily on the fact that these two species have, on the average, two less vertebrae (one cervical, one sacral) than do B. bernicla and B. canadensis. Verheyen also neglected to designate a type species for “Eubranta” . Even if the segregation of these two species as a separate genus from Branta were warranted, which I do not believe, the introduction of a new generic name was completely unnecessary. Both of thse species, according to the synonymies presented by Hellmayr and Conover (1948, p. 294-295), have been named as monotypes of new genera; Leucopareia Reichenbach, 1852, for leucopsis, and both Rufibrenta Bona- parte, 1856, and Ptocas Heine, 1890, for ruficollis. In order to allocate the name Eubranta Verheyen, 1955, I here designate its type as Anas leucopsis Bechstein, and thus place it as an absolute synonym of Leucopareia Reichen- bach, 1852, a genus currently believed inseparable from Branta Scopoli, 1769. A brief distributional note on the Canada Goose may well be placed here. Hanson and Smith (1950, p. 76-77) believed southern New Jersey to be the northernmost part of the Atlantic coast reached by migrating and wintering Branta canadensis interior Todd. The A.O.U. Check-list (1957, p. 61) goes even further, stating “Not recorded from the Atlantic coast north of Mary- land”. However, I have examined three specimens of this race from Montauk Point, Long Island, N. Y., in the collection of the American Museum of Natural History (A.M.N.H. 350131, December 3, 1909; A.M.N.H. 350133 and 350134, March 14, 1902). Hellmayr and Conover (1948, p. 305) listed a specimen in the Chicago Natural History Museum from Rockaway Beach, Queens County, New York. 3. The genus Chen The status of the Blue and Snow geese, long one of the most difficult questions in systematic ornithology, has been under investigation for some 120 Annals of Carnegie Museum vol. 35 years by Graham Cooch, and I do not venture to discuss the matter here. I can not, however, refrain from stating that I can see no justification for the continued recognition of the Blue Goose, Chen caerulescens, as a sepa- rate species in the A.O.U. Check-list (1957). Whether the relationship of the Blue and Lesser Snow geese is best expressed by calling them subspecies or color phases will be determined by specialists in this group, but to call these two forms separate species is to ignore half a century of progress in the study of evolution. Many modern authors include Chen in an expanded genus Anser. This is entirely a matter of subjective preference as to size and scope of genera, since the two are undeniably closely related. 4. Anser albifrons Delacour (1954, p. 110, and personal communication) has stated that all records known to him of presently existing specimens of White-fronted Geese along the east coast of North America pertain to the Greenland race, Anser albifrons flavirostris Dalgety and Scott. One of the most striking characteristics of this race is the possession of an orange-yellow rather than a pink bill. Stoner (1944) published the details of a specimen killed near Rouses Point, Lake Champlain, N. Y., on October 22, 1943. Although the specimen was not preserved, both measurements and color notes were taken. The description of the bill as “pink” precludes the identification of this specimen as flavirostris . The measurements indicate that it was neither the small European A . a. albifrons nor the very large far western A. a. gambelli. This leaves, by elimination, A. a. frontalis , the common White-fronted Goose of western North America, which migrates chiefly west of the Mis- sississippi. An occasional eastern stray of this form would hardly be sur- prising. I have examined the series of European specimens of Anser albifrons men- tioned by Todd (1950, p. 64). Comparing four from Holland with eight from Austria, Montenegro and Albania, Todd wrote that the former “differ in the darker, browner coloration of their upper parts and wings and in the more brownish suffusion of the neck and under parts generally. The significance of this variation I do not presume to explain beyond suggest- ing that, since it cannot be seasonal, it could be racial.” Although not directly pertinent to a North American bird, this matter may appropriately be settled in the present discussion of the species involved. The color differ- ences between the two series noted by Mr. Todd may be easily explained, and have nothing to do with geographic variation. The Holland birds were collected in 1892 and 1900, and are foxed and stained. The others were collected in 1929 and 1932, are clean specimens, and have never been on exhibition as the Holland birds were. 5. The Mallard and its relatives In an earlier paper (Parkes, 1954, p. 152) I commented on a statement made by Delacour and Mayr (1945, p. 21) who wrote that “it seems obvious that the Mexican and Black Ducks ( diazi and rubripes) are only sub- specifically distinct from the Dusky Duck ( fulvigula ).” Delacour and Mayr united these three forms under fulvigula, the oldest name. My comments were as follows: “It is my belief that the case is by no means so ‘obvious'. 1958 Parkes: Systematic Notes on North American Birds 121 The Mexican Duck, Anas diazi, is so close to the Mallard, A. platyrhynchos, that a case might be made for considering it a rather restricted ‘hen-feathered' subspecies of Mallard, except that diazi and platyrhynchos are sympat- ric (Lindsey, 1946, p. 484). A comparatively recent origin of diazi from platyrhynchos is suggested by the high frequency of hybridization (Lindsey, 1946, p. 484).” I then went on to point out that the Dusky Duck, A fulvi- gula, is in many respects about midway between the Mallard and the Black Duck. Delacour himself apparently came to doubt the “obvious” conspecificity of rubripes, diazi and fulvigula. In his recent book (Delacour, 1956) he unites the latter two forms with platyrhynchos , allowing rubripes to stand as a full species. He gives no references to support this treatment, nor does he explain his own change of mind. The range map (on page 41 of his book) does not show the overlap of the breeding ranges of diazi and platyrhynchos , He states that “They [diazi] do not seem to mix with wintering Mallards which are often found at the same localities during the winter”, a state- ment completely at variance with the New Mexico observations of Lindsey (1946). He mentions the fact that drakes of diazi “sometimes have more or less curled up central tail feathers and traces of bright colours . . . thus showing a close relationship to the common Mallard.” He gives no indica- tion as to whether such birds are ever found outside the area of overlap of diazi and platyrhynchos , nor does he even mention the hybridization de- scribed by Lindsey and mentioned in the A.O.U. Check-list (1957, p. 72). The treatment of the Dusky Duck ( fulvigula ), now called Florida Duck by Delacour, is equally scanty and devoid of explanation. At present I can see no compelling reason to alter my statement of 1954 that “All in all, I prefer to consider the Mallard, Black Duck, Dusky Duck and Mexican Duck as specific entities.” 6. The Green-winged Teal The conservatism of the A.O.U. Check-list is nowhere illustrated better than by its persisting in giving full specific rank to the American Green- winged Teal (Anas carolinensis). This provincial viewpoint has been aban- doned by virtually all students of waterfowl the world around. As is well known, the females of the American and the European (A. crecca ) Green- winged Teal are virtually indistinguishable. The two are geographic repre- sentatives, but individual birds of each of the races occasionally stray within the range of the other. Hybridization among ducks is so common, of course, that it can not be used as a sole criterion of conspecificity, but it is inter- esting to note that Cruickshank (1986) and Poole (1940) have described apparent hybrids or intergrades between carolinensis and crecca. (See Parkes, 1955b, p. 38, for further discussion of this case.) There are no behavioral characters, often useful in duck classification, to separate the two forms. The American Green-winged Teal should be known as Anas crecca carolinensis Gmelin. 7. The Shovellers Delacour and Mayr (1945, p. 17) and Delacour (1956, p. 19) have re- emphasized the extremely close relationship among the four shovellers (“Spatula”) and the three blue-winged “teal” (“Querquedula”). They make 122 Annals of Carnegie Museum vol. 35 the interesting and, to me, highly plausible suggestion that “the shoveller group is polyphyletic, owing its origin to the repeated development of large-sized and large-billed species from the original blue-winged duck stock” (Delacour and Mayr, 1945, p. 17). Oliver (1954, p. 193) objected to this idea, with its necessary corollary of combining “Spatula” with Anas (in- cluding “ Querquedula” ). Oliver’s full statement is as follows: “Spatula differs from Anas in important bill characters and consists of closely related species occupying different continents. This shows [italics mine] that the species have not risen independently in each continent from different species of Anas . Spatula is so different from Anas that its union with that genus would cover up an important morphological characteristic and make Anas indefinable. It certainly should be kept as a genus distinct from Anas.” This treatment is adopted by the conservative A.O.U. Check-list (1957), which interposes the widgeons Mareca ”) between the Shoveller and its closest relatives, the Blue-winged Teal and Cinnamon Teal, included in Anas by the A.O.U. Check-list. Meinertzhagen (1951, p. 444) pursues the concept of monophyletic origin of the shovellers to the ultimate extreme, and makes all of them subspecies of Spatula clypeata! As amply shown by Delacour and Mayr, the shovellers and the blue- winged “teal”, taken together, constitute a well-knit group within the ex- panded genus Anas. Extreme variation in bill size notwithstanding, these ducks are obviously more closely related to one another than any is to the rest of the genus Anas. This is supported by plumage pattern, feeding habits, courtship display, and, to some extent, voice. Two wild-taken hybrids be- tween Anas (“ Spatula ”) clypeata and A. (“ Querquedula ”) discors have been described by Childs (1952). I saw what appeared to be such a bird myself at the Montezuma Federal Waterfowl Refuge, in central New York, on October 11, 1952. Delacour (1956, p. 182-183) makes the interesting point that such hybrids are extremely similar in appearance to the Australian Shoveller (A. rhynchotis). In order, then, to justify retention of the genus Spatula for the four shovellers, Oliver (and presumably the A.O.U. Check-list) must rely entirely on the shape of the bill, a notoriously unreliable character in avian taxonomy at generic and higher levels. The geographic distribution of the shovellers (one holarctic, three in the Southern Hemisphere) and the remarkable simi- larity in color and plumage between the Cinnamon Teal ( cyanoptera ) and the South American Shoveller (platalea), and between the Blue-winged Teal (discors) and the Australian Shoveller (rhynchotis), are strong evidence in favor of independent origin of shovellers from blue-winged duck stock in at least three different cases. Such a polyphyletic origin would, of course, preclude segregation of the large-billed forms as a genus Spatula , much less combining them all as subspecies of Spatula clypeata. The Cape Shoveller (Anas smithi = Spatula capensis of authors) is a somewhat different case. This African species is virtually a “hen-feathered” version of the holarctic A. clypeata, and may well be derived from the latter species, which migrates to Africa. These two could be considered to constitute a superspecies; the “shoveller” bill would then have been independently derived only three times. 1958 Parkes: Systematic Notes on North American Birds 12S 8. The Greater Scaup Witherby et al. (1939, p. 308), Scott (1949, caption to plate XV), and some other authors have cast doubt on the validity of the New World sub- species of the Greater Scaup, Aythya marila nearctica Stejneger. With this in mind I examined the extensive series of this species in the American Museum of Natural History, and found that nearctica is readily recognizable by the coarse black barring of the upper parts, exactly as characterized by Hellmayr and Conover (1948, p. 371, footnote). The geographically inter- mediate A. m. mariloides (Vigors) of eastern Asia is somewhat intermediate in color between marila and nearctica, but is smaller than either. (For measurements, see Hartert, 1920, p. 1344.) 9. The Spectacled Eider The fifth edition of the A.O.U. Check-list (1957, p. 91) places the Spec- tacled Eider in the genus “Lampronetta” . In the thirty-first supplement to the fourth edition of the Check-list (Wetmore et al., 1956, p. 448), the pro- posed change from Arctonetta Gray, 1856, as used in the fourth edition, to Lampronetta Brandt, 1847, was announced. The reference cited for this change was a paper of mine (Parkes, 1955a). It is true that in this paper I pointed out that Lampronetta antedated Arctonetta. However, my wording was intended to make it plain that I was certainly not advocating the con- tinued recognition of a monotypic genus for the Spectacled Eider under any name; I pointed out that this species possesses no trenchant characters to separate it from Somateria, and mentioned the priority of Lampronetta over Arctonetta to indicate that a name change for this species was inevitable in any case. Lest the citation of my paper in connection with the change in A.O.U. usage be misleading, I wish to reiterate my firm belief that the Spectacled Eider can not be separated from the genus Somateria . REFERENCES American Ornithologists’ Union 1957. Check-list of North American birds, ed. 5. 691 p. Childs, Henry E., Jr. 1952. Hybrid between a Shoveller and a Blue-winged Teal. Condor, v. 54, p. 67-68. Cruickshank, Allan D. 1936. Some observations of the European Teal. Auk, v. 53, p. 321*322. Delacour, Jean 1954. The waterfowl of the world, v. 1, 284 p. London. 1956. The same. v. 2, 232 p. London. Delacour, Jean, and Ernst Mayr 1945. The family Anatidae. Wilson Bulletin, v. 57, p. 1-55. Hanson, Harold C., and Robert H. Smith 1950. Canada geese of the Mississippi Fly way. Bulletin of the Illinois Natural History Survey, v. 25, p. 67-210. Hartert, Ernst 1920. Die Vogel der palaarktischen Fauna, v. 2, pt. 4, p. 1217-1344. 124 Annals of Carnegie Museum vol. 35 Hellmayr, Charles E., and Boardman Conover 1948. Catalogue of birds of the Americas and the adjacent islands. Field Museum of Natural History Zoological Series, v. 13, pt. 1, no. 2. 434 p. Lindsey, Alton A. 1946. The nesting of the New Mexican Duck. Auk, v. 63, p. 483-492. Meinertzhagen, R. 1951. Some relationships between African, Oriental and Palaearctic genera and species, with a review of the genus Monticola. Ibis, v. 93, p. 443-459. Oliver, W. R. B. 1954. Avian taxonomy. Emu, v. 54, p. 190-198. Parkes, Kenneth C. 1954. Notes on some birds of the Adirondack and Catskill moun- tains, New York. Annals of the Carnegie Museum, v. 33, p. 149-178. 1955a. The generic name of the Spectacled Eider. Auk. v. 72, p. 85-86. 1955b. Sympatry, allopatry, and the subspecies in birds. Systematic Zoology, v. 4, p. 35-40. 1955c. Systematic notes on North American birds. 1. The herons and ibises (Ciconiiformes). Annals of the Carnegie Museum, v. 33, p. 287-293. Peters, James L. 1931. Check-list of birds of the world, v. 1, 345 p. Cambridge, Mass. Poole, Earl L. 1940. Recent records from Lake Ontelaunee, Pennsylvania. Auk. v. 57, p. 577-578. Schiller, E. L. 1925. Danmarks Fugle, v. 1, 552 p. Copenhagen. Scott, Peter 1949. Key to the wildfowl of the world. Severn Wildfowl Trust Second Annual Report, p. 91-137. Simpson, George G. 1945. The principles of classification and a classification of mammals. Bulletin of the American Museum of Natural History, v. 85, 350 p. Stoner, Dayton 1944. White-fronted Goose at Rouses Point, New York. Auk, v. 61, p. 651-652. Todd, W. E. Clyde 1950. Nomenclature of the White-fronted Goose. Condor, v. 52, p. 63-68. Verheyen, Rene 1955a. La syst&natique des Anseriformes bas^e sur l'ost^ologie com- pare (suite). Bulletin Institut Royal des Sciences Naturelles de Belgique, v. 31, no. 36, p. 1-16. 1955b. The same (fin). Bulletin Institut Royal des Sciences Naturelles de Belgique, v. 31, no. 38, p. 1-16. 1958 Parkes: Systematic Notes on North American Birds 125 Wetmore, Alexander 1951. Observations on the genera of the swans. Journal of the Wash- ington Academy of Sciences, v. 41, p. 338-340. Wetmore, Alexander (Chairman) and others 1956. Thirty-first supplement to the American Ornithologists’ Union Check-list of North American birds. Auk, v. 73, p. 447-449. Witherby, H. F., F. C. R. Jourdain, N. F. Ticehurst and E. W. Tucker 1939. The handbook of British birds, v. 3, 387 p. London. Yamashina, Yoshimaro 1952. Classification of the Anatidae based on the cyto-genetics. Papers from the Coordinating Committee for Research in Genetics, v. 3, p. 1-34. y - PUf^ ART. 9. A RECENT FISSURE BEDFORD COUNTY, PENNS By John E. Guilday* and Martin This paper is a report on a collection of vertebrate skeletal material re- covered from the New Paris sink-holes. These fissures are located 1.5 miles northeast of New Paris, West St. Clair Township, Bedford County, Pa., on the property of Oscar Miller. Fig. 1. The bone deposits were discovered by J. Howard Taylor of New Paris, Pa., who broke into the top of fissure (sink-hole) no. 2, then evidenced by a slight depression on the surface of the ground. Mr. Taylor made a small collection of elk bones which included the most interesting single specimen from the fissures, two elk vertebrae with an imbedded flint arrowhead. The locality was visited in February, 1932, by Charles E. Mohr, and reported on by him in Stone’s ‘ Pennsylvania Caves” (ed. 2, 1932) and ‘‘Descriptions of Pennsylvania’s Undeveloped Caves” (1953). In April, 1948 a group com- posed of members of the Pittsburgh Explorers Club and the Pittsburgh Grotto, National Speleological Society, visited the area. Mr. Taylor led the group and volunteered valuable information about the site. As a result of this trip the major part of an elk skeleton was excavated and donated to Carnegie Museum. With the aid of a grant from the A. W. Mellon Educa- tional and Charitable Trust, the Museum was able to institute field work at the site. Excavations were carried out during 1949 and 1950 under the direction of Dr. J. Leroy Kay. Actual field work was under the direction of Albert C. Lloyd. Oscar Miller not only granted the Museum permission to work at the site but also served as an active and enthusiastic member of the excavating party. The project is indebted to the Pittsburgh Grotto, National Speleological Society, whose members have put so much time and effort on it. Special thanks are due Albert H. Bauer, Ralph C. Bossart and John A. Leppla of Pittsburgh for their assistance in the excavation. The writers would like to express their gratitude to Dr. J. Leroy Kay and Dr. J. Kenneth Doutt for the privilege of studying the collection and for much helpful advice and assistance; to Neil D. Richmond, Curator of Am- phibians and Reptiles, Carnegie Museum, for examining the snake remains; to George B. Thorp and Miss Lillian Heeren for the surveying and drafting of the map, respectively; and to Miss Caroline A. Heppenstall, Assistant Curator of Mammals, and Mrs. Alice M. Guilday for their assistance in the project. A popular article concerning the fissures appeared in Carnegie Magazine (Guilday, 1948) and in 1955 a report upon the carnivores of the New Paris sink-holes by the junior author was submitted to the University of Pitts- burgh in partial fulfillment of the requirements of a Master of Science degree, and is on file at the University of Pittsburgh library (Bender, [1955]). The New Paris sink-holes treated in this report are a group of vertical limestone caverns. They occur in the Helderburg (lower Devonian) Lime- stone. It is not obvious whether the sink-holes are solution enlargements of * Assistant Curator of Comparative Anatomy, Carnegie Museum. •[Instructor in Geology, University of Pittsburgh. Submitted for publication, November 5, 1957. Issued November 20, 1958. 127 3ivn i institution 1 1958 128 Annals of Carnegie Museum vol. 35 structural rock joints or whether they were positioned by other factors, but the former seems more probable. The Helderburg Limestone is exposed on the Schellsburg anticline (locally known as Chestnut Ridge) as a series of inliers which occupy the valley areas on the flanks of the anticline with low ridges between them. These inliers are surrounded by the outcropping Oriskany formations that form the high ground of the anticline. Physiographically the New Paris sink-holes lie within the Ridge and Valley Province, just east of Allegheny Mountain. This mountain lies at the eastern boundary of the Allegheny Plateau. They are on the west flank of Chestnut Ridge, at an altitude of 1500-1600 feet. The immediate area of the fissures is well drained and covered with hardwoods. The surrounding areas are under cultivation. Biologically the area lies within a zone of transition be- tween the Carolinian Biotic Province and the more northerly Canadian Biotic Province. (Dice, 1943). Gifford and Whitebread (1951) state that the Ridge and Valley section generally is difficult to assign to either of these two broad classes. They state (p. 18), “No groups of mammals were found that could be classed as either northern or southern. For the most part the mammals that occur together in any one locality here are determined by the availability of suitable habitats and the adaptability of the species.” The collection includes samples from three sink-holes, all geologically related but not in direct contact (See Fig. 1). Sink-hole no. 4 or “Lloyd’s rock hole” is currently being excavated by the Pittsburgh Grotto, National Speleological Society, but, to date, the work has produced no fossil fauna.* Sink-hole no. 1 or “ Kantner’ s Sink”. Known as the “Round sink-hole”. Excavated by the Museum party, this sink-hole proved unproductive and received only cursory attention. A small collection of small mammal bones was made. Sink-hole no. 2 or the “ Elk hole”. This was the original bone deposit discovered by Mr. Taylor and it was completely excavated by the Museum party. This sink-hole was not open to the surface at the time of discovery but was sealed by a cover of tree roots, dirt and organic debris. This fissure had an initial depth of 47 feet. Excava- tions were carried to a further depth of 12 feet at which point the bone de- posit was flooded by ground-water and operations ceased. The bulk of this report concerns the collection from this fissure. Sink-hole no. 3. A small collection of both large and small mammal bones from this fissure was donated by members of the Pittsburgh Grotto, National Speleo- logical Society, and the Pittsburgh Explorers Club. This was augmented by collections made by Arnold D. Lewis, Allen D. McCrady and the senior author of this paper, prior to the Museum excavations. This material was picked up on or close to the surface of the rock debris at the bottom of the fissure in two areas— one at the bottom of the first landing (approximately 50 feet) and the other at the bottom of the sink-hole (87 feet). The con- temporaneity of these two deposits was indicated by the finding of frag- #Since the above was written, an extensive Pleistocene fauna has been recovered from sink-hole no. 4, and is currently under study. 1958 Guilday and Bender: Fissure Deposit in Pennsylvania 129 Fig. 1. Maps showing location of New Paris sink-holes Main map gives detailed compass directions for individual sink-holes. Insert map gives general location of sink-holes described in this article. To reach the point marked Miller house, proceed north on Route 96, 0.3 mile from intersection in center of New Paris, take right fork (NE.) for 0.3 mile, turn right (SE.) 0.1 mile, turn left (NE.), continue straight 0.45 mile to Miller house. 130 Annals of Carnegie Museum vol. 35 ments of the same black bear skull at both levels. This fissure was not ex- cavated and more skeletal material remains but any extensive digging would produce a rock slide. The bone-bearing deposit at the bottom of the fissures was a heterogeneous mixture of clay, apparently resulting from chemical breakdown of the lime- stone walls, angular fragments of limestone varying in size from pebbles to blocks many tons in weight, organic debris and bones. No evidence of selec- tive sorting by water was noted and no evidence of internal stratification was observed in the deposits. Apparently the animals blundered into the holes over a period of many years; were either killed by the fall or starved to death and the deposit grew, augmented through the years by falls of rock, soil and additional victims. Oscar Miller, the owner of the property, states that the fissures of this area have been alternately opened and closed to the surface by varying accidents of forest litter accumulation and periodical cave-ins. The bones were, for the most part, unbroken, well preserved, and in the case of some of the skeletons from sink-hole no. 2 still lay in attitudes of articulation. Some of the bones from this fissure were spattered with a thin calcareous patina, but none of the bones from the fissures showed any noticeable degree of mineral replacement. The numbers of bones listed below refer only to items which were readily diagnostic; such as limb bones, skulls, and teeth, and on these were based the estimates for the minimum number of individuals represented. No attempt was made to list the thousands of minor skeletal parts present, except in special cases. To attempt to do so would have proved both futile and impractical. The assumption is made that a given animal was trapped in its entirety and, for census purposes, a count of skull elements would give a reliable estimate. All bones were carefully examined however. The bones recovered from the fissure deposits represent species all living today, although several do not inhabit the area at present. The elk and the wolf were exterminated in Pennsylvania in the late 1800’s (Rhoads, 1903). A flint arrowhead, embedded in two coalesced cervical vertebrae of an elk, places the age of the deposit in Post-Wisconsin times, most probably well into the Recent and possibly as late as the eighteenth century. Absence of re- mains of the red fox and the opossum (from sink-hole no. 2), two animals now present in the area but believed to have been absent during late pre- historic times, suggests the possibility that the deposit predates the period of profound ecological changes ushered in by European colonization. Some of the smaller mammals have been trapped since the most recent opening of the fissure. This is very apparent in the fresh condition of some of the rabbit remains. A woodchuck fell in during the excavating operations. The bulk of the collection appears to be of some historical antiquity, although geologically recent. There are 67 forms of mammals representing 47 genera recorded from Bedford County within recent times. (Gifford and Whitebreacl, 1951). By comparison, 26 species of 25 genera were found in the fissure. While the animals recovered from the fissures in no way represent an adequate cross-section of the total mammalian fauna within the area during historic times, they clo offer important clues to the environment of that area during the time the fissure traps were in operation. Remains of such 1958 Guilday and Bender: Fissure Deposit in Pennsylvania 131 forms as the flying squirrel, red squirrel, gray squirrel and porcupine, and the high numbers of chipmunk and pine mouse remains (See discussion under Pitymys) all point directly to a former forest cover. The total fauna is one that would be expected in the area at the present day, with the obvious exception of those forms which were exterminated by man. In this instance the forested nature of the country during the recent past is well known and the fact comes as no great surprise, but it does serve to strengthen the reliability of fossil faunas as indicators of past climatic and ecologic condi- tions, provided the habitat requirements of the component species are well enough known. ANNOTATED LIST OF VERTEBRATE REMAINS (Unless otherwise noted, all specimens are from sink-hole no. 2) amphibians Remains of three large toads, Bufo sp., represented by tibiotarsi and skull fragments. These agreed closely with specimens of Bufo americanus used for comparison and are probably referable to that species. Two tibiotarsi were referred to the genus Rana. The wood frog Rana sylvatica was found living in the sink-hole and the surrounding woodland. reptiles Order Chelonia Family Emydidae Box turtle. Terrapene Carolina (Linnaeus) Remains of four individuals. Shell fragments of one individual were re- covered from sink-hole no. 3. Order Serpentes Family Colubridae Black rat snake. Elaphe obsoleta Say One right maxilla, one left mandible and vertebra could not be distin- guished from those of a 52-inch female black rat snake collected from the fissure area during the course of the field work. Family Crotalidae Timber rattlesnake. Crotalus horridus Linnaeus At least three individuals. Represented by skull elements which include three right and one left maxilla, mandibles and numerous vertebrae. The unidentified material, largely vertebrae, was kindly examined by Neil D. Richmond who reports as follows: Most of the snake vertebrae are Crotalidae. These appear to be of two types — a light, more delicate type and a large massive type that are referable to Crotalus. The maxilla and other bones of the skull are definitely referable to Crotalus. Whether the more delicate vertebrae are young Crotalus or are those of Ancistrodon (copperhead) can not be determined with the material at hand. The most numerous Colubrid vertebrae are those of Coluber. The vertebrae from the sinks agree in minute detail with Pennsylvania Coluber constrictor skeletons. One lot of several vertebrae agrees with Elaphe. Two lots of vertebrae, apparently of quite different ages as one is discolored by iron salts, agree with Heterodon (hog-nosed snake). Several small Colubrid vertebrae have been left unidentified as they were not found together and are probably anterior thoracic vertebrae of one of the other Colubrid snakes. 132 Annals of Carnegie Museum vol. 35 BIRDS Order Strigiformes Family Strigidae Great horned owl. Bubo virginianus One partial skeleton. Unidentified. Partial skeleton of one small perching bird. MAMMALS Order Marsupialia Family Didelphiidae Opossum. Didelphis marsupialis virgin iana Kerr No opossum remains were recovered from sink-hole no. 2. One skull and one cervical vertebra of an adult opossum were picked up from the super- ficial deposits in sink-hole no. 3. It should be noted that the latter fissure had been open to the surface at the time of discovery. The opossum is a recent addition to the fauna of Pennsylvania. With the opening up of the country and the removal of much of the primeval forest the opossum spread to the north. It is a common animal in Bedford County today but was rare or absent during prehistoric times. Parker (1953) reports the finding of an opossum “in an advanced state of purification” in one of the fissures sub- sequent to the Museum excavations. Order Insectivora Family Soricidae Long-tailed shrew. Sorex sp.? At least three individuals. Short-tailed shrew. Blarina brevicauda (Say) At least 31 individuals represented by 74 skull and mandible fragments. Family Talpidae Hairy-tailed mole. Parascalops breweri (Bachman) One right clavicle and one right humerus. Order Chiroptera Family Vespertilionidae Little brown bat. Myotis sp.? At least three individuals. One bat skull, without mandibles, from sink-hole no. 3, was identified by Dr. J. Kenneth Doutt as Myotis , probably lucifugus. Bat. species? At least two individuals. Order Lagomorpha Family Leporidae Cottontail rabbit. Sylvilagus floridanus (J. A. Allen) Two complete skulls. In addition, one complete skull from sink-hole no. 3; this appeared to be extremely recent in origin compared with other bones from the same fissure. New England cottontail rabbit. Sylvilagus transitionalis (Bangs) One complete skull and one pair of frontal bones seem to be referable to this species. Both are of adult animals and in both the supraorbital processes agree in size and shape with those of the transitionalis skulls used for com- parison. In addition, two complete skulls from sink-hole no. 3 are referred to this species. 1958 Guilday and Bender: Fissure Deposit in Pennsylvania 133 Cottontail rabbit. Sylvilagus (either floridanus or transitionalis ) At least 13 individuals represented by 48 skull and jaw fragments, plus 87 postcranial bones and fragments. Seven mandibles, five skull fragments, and 34 postcranial bones from sink-hole no. 3. There is some evidence that S. transitionalis was the only form of the cot- tontail rabbit present in Bedford County during prehistoric times (Rhoads, 1903). Subsequent destruction of suitable habitat within the area has reduced its former range. The New England cottontail is now rare and local in Bed- ford County (Gifford and Whitebread, 1951, p. 31). Order Rodentia Family Sciuridae Woodchuck. Marmota monax (Linnaeus) At least eight individuals represented by four skulls, 13 disassociated lower jaws and 23 postcranial bones. During the period of excavation a wood- chuck was found dead at the bottom of the fissure. It is preserved as a study specimen in the Section of Mammals, Carnegie Museum. Two mandi- bles and seven postcranial bones from sink-hole no. 3. Chipmunk. Tamias striatus (Linnaeus) At least 109 individuals, represented by 139 skulls and skull fragments, 215 mandibles and 553 postcranial bones. This was the commonest mammal encountered in the deposit. One right mandible was recovered from sink- hole no. 3. Gray squirrel. Sciurus carolinensis Gmelin Two skulls are certainly this species. Four skull fragments, six mandibles and 15 postcranial bones are referred on the basis of size. A total of four in- dividuals is represented. Red squirrel. Tamiasciurus hudsonicus (Erxleben) One complete skull from sink-hole no. 3. Flying squirrel. Glaucomys volans (Linnaeus) At least 10 individuals, represented by 27 skull and mandible fragments and 50 postcranial bones. It is conceivable that the northern flying squirrel, Glaucomys sabrinus (Shaw), could be represented as well in this collection. It has been taken as close as 56+ miles at McGees Mills, Clearfield County, Pa. (Roslund, 1951). All of the flying squirrel bones from the fissure agree in size with specimens of Glaucomys volans, but the skeletal differences be- tween the two species are too subtle to permit an accurate specific identifica- tion of the more or less fragmentary material. Glaucomys sabrinus has not been taken in the Ridge and Valley Section of Pennsylvania to date (Gifford and Whitebread, 1951). Family Cricetidae White-footed mouse. Peromyscus sp.? At least 28 individuals represented by 74 fragments of skull and mandibles. The species is undoubtedly either Peromyscus leucopus (Rafinesque) or Peromyscus maniculatus (Wagner). An additional mandible was recovered from sink-hole no. 3. Lemming vole. Synaptomys cooperi Baird One individual, represented by a skull and associated lower jaws. Meadow vole. Microtus pennsylvanicus (Ord) Three individuals represented by three skulls. 134 Annals of Carnegie Museum vol. 35 Pine vole. Pitymys pinetorum (LeConte) Fifty-one individuals, represented by 48 skulls, four additional skull frag- ments and three mandibles. Remains of small mammals are generally assumed to be more reliable indicators of past environmental conditions than are those of the more widely ranging larger mammals (Hibbard, 1955, p. 200). The high percent- age of pine vole remains and the low percentage of meadow voles is inter- esting. The pine vole is much more of a burrowing animal than is the meadow vole and appears to be limited to areas of loose friable soils irre- spective of the type of ground cover. In the mountainous areas of Pennsyl- vania it is local in distribution and more commonly met with in heavily wooded areas than is the meadow vole which prefers a more open meadow habitat and is rarely taken by the collector in wooded areas. These require- ments are approximations of optimum habitats and in periods of high vole abundance may have lessening significance. The ratio of 17:1 (51 Pitymys , 3 Microtus) in favor of the pine vole reflects the forested nature of the area during the time the sink-holes were open to the surface. Both of these rodents are present in Bedford County today, Microtus being the more common, but the amount of woodland is drastically different now from what it was in prehistoric times. In 1945 only 35.8% of Bedford County was classed as forested (Gifford and Whitebread, 1951). ' Meadow mouse or pine mouse. Microtinae sp.? Ninety-six mandibles and three skull fragments were either too fragmen- tary or had lost too many teeth to permit a satisfactory identification closer than subfamily. Most of the mandibles undoubtedly belong to the skulls listed above as Pitymys pinetorum. One Microtine mandible was recovered from sink-hole no. 3. Muskrat. Ondatra zibethicus (Linnaeus) Remains of at least one individual, represented by one skull and four postcranial bones from sink-hole no. 3. Muskrats wander widely at times. The hilly nature of the sink-hole area is and was entirely unsuitable as muskrat range. This was undoubtedly a chance inclusion of a stray animal and has no bearing upon the past en- vironment of the fissure area. Family Erelhizontidae Porcupine. Erethizon dorsatum (Linnaeus) At least two individuals represented by one skull, four mandibles and 12 postcranial bones. At least five individuals represented by two skulls, seven mandibles (two left, five right), 10 isolated teeth and 13 postcranial bones from sink-hole no. 3. Order Carnivora Family Canidae Timber wolf. Canis lupus lycaon Schreber) Two individuals, both adult but varying somewhat in size, were recovered. The bones were badly shattered and the remains of the two wolves were intermhigled. The fragmentary skeletons were found immediately beneath an articulated, undamaged elk skeleton. 1958 Guilday and Bender: Fissure Deposit in Pennsylvania 135 Gray fox. Urocyon cinereoargenteus (Schreber) One adult and four pups, possibly litter mates. Represented by two skulls, two skull fragments, eight mandibles and eight postcranial bones. At least one individual (adult) represented by six fragments, from sink-hole no. 3. Family Ursidae Black bear. Euarctos americanus (Pallas) One partial skull and associated teeth. This was a young animal, still re- taining the milk dentition and with open sutures. In sink-hole no. 3 a right mandible, a right maxilla, and one right humerus of a large and, judging from the amount of tooth wear, very old black bear. Family Procyonidae Raccoon. Procyon lotor (Linnaeus) One upper molar and one partial right mandible of an immature animal. Family Mustelidae New York weasel. Mustela frenata Lichtenstein One skull and one left mandible. Judging from its large size this specimen was a male. Striped skunk. Mephitis mephitis (Schreber) At least five individuals (four adult and one juvenile) represented by four skulls, five mandibles and 1 1 postcranial bones. In addition, three postcranial bones from sink-hole no. 3. ) Order Artiodactyla Family Cervidae Elk. Cervus canadensis canadensis (Erxleben) Two articulated skeletons were found. The posterior half of a skull bear- ing the antler bases, several segments of antler, a humerus and two cervical vertebrae were found by J. Howard Taylor of New Paris, Pa., during his original investigations of the fissure. The remaining bones of the same skele- ton were recovered by the members of the Pittsburgh Grotto, National Speleological Society. Both collections are now at Carnegie Museum. This was an adult male. Although the antlers were fragmentary, enough remains to show they were of some size and were firmly attached to the skull at the time of death. The antlers were broken from the skull at a point out on the antlers themselves, not at the pedicel where natural shedding occurs. With the exception of the antlers, which were badly rotted, the skeleton was in an excellent state of preservation. The following bones were missing: pre- maxillae; left maxilla and tooth row; nasals; palate; zygoma; cervical verte- brae, numbers one, two and four; thoracic vertebrae, number one; caudal vertebrae, all missing; 12 phalanges; all “dew claws”; ribs number 11, 12 and 13; two fibulae; right metatarsal; right patella; left radius; left meta- carpal; all carpals and hyoids. Cervical vertebrae numbers six and seven were completely coalesced as a result of regenerative processes involved in the healing of what must have been a severe arrow wound. The flint arrowhead, with its base broken off, was embedded at the base of the left transverse process of the sixth cervical. The animal aparently survived and the wound had completely healed. The arrowhead was so extensively enveloped in new bone tissue that it can not be removed without destroying some of the ver- tebra itself. An x-ray of the specimen, taken by Dr. J. Kenneth Doutt and 136 Annals of Carnegie Museum vol. 35 Dr. E. G. Meisel, yielded no additional information. Mr. Taylor is to be congratulated for his efforts to preserve this unique and important specimen as well as for his generosity in presenting it to Carnegie Museum. The specimen is illustrated in Fig. 2. The second skeleton, that of a doe, was found by A. C. Lloyd of the Museum field party. It was a more nearly complete skeleton than that of the bull. Its skull, protected from crushing by the surrounding rocks, was recovered intact. Elk remains were also recovered from sink-hole no. 3. These were dis- articulated, represented parts of more than one individual and in some cases were badly shattered. They include the following: 4 upper molars, 3 upper premolars, sacrum, 13 vertebrae, 7 ribs, 2 sternal ribs, 3 sternebrae, 1 pelvis, 2 femora, 2 tibiae, 2 metatarsals, 2 humeri, 2 radii, 2 ulnae, various carpals, tarsals and phalanges. Also a large number of bones of a very young elk fawn, including a complete but disarticulated skull. White-tailed deer. Odocoileus virginianus (Boddaert) Remains of three individuals, presumably two antlered bucks and one fawn. Represented by five skull fragments, three broken antlers and 14 post- cranial bones. In addition, a right tibia, a patella, fragmentary scapula, teeth and sacrum of apparently one adult, as well as one right mandible of a fawn were recovered from sink-hole no. 3. There is ample evidence in Pennsylvania and neighboring states for cave deposits dating from the Pleistocene (Hay, 1923). While the fauna recovered to date from the New Paris fissures is recent and of no geological antiquity, it is possible that further prospecting along Chestnut Ridge (Bedford County) would be profitable. The area is rife with fissures and it is possible that some of them might contain the fossil remains of Pleistocene mammals. REFERENCES Bender, Martin S. [1955.] The carnivores of the New Paris sink-holes. Ill p. Master’s thesis at University of Pittsburgh. Dice, Lee R. 1943. The biotic provinces of North America. University of Michi- gan Press, Ann Arbor, 78 p. Gifford, Clay L., and Ralph Whitebread 1951. Mammal survey of south central Pennsylvania. Final Report Pittman-Robertson Project 38-R. 75 p. Pennsylvania Game Commission, Harrisburg, Pa. Guilday, John E. 1948. The fallen monarch. Carnegie Magazine, v. 22, no. 2, p. 46-49. Hay, Oliver P. 1923. The pleistocene of North America and its vertebrated animals from the states east of the Mississippi River and from the Ca- nadian Provinces east of longitude 95°. Carnegie Institution of Washington. 499 p. (Publication no. 322.) 1958 Guilday and Bender: Fissure Deposit in Pennsylvania 137 Fig. 2. Sixth and seventh cervical vertebrae of an adult male elk ( Cervus canadensis ), showing embedded arrowhead A. Spinous process, 7th cervical. B, B\ Demifacets for the 1st ribs. C. Centrum of 7th cervical. D, D'. Postzygapophyses, 7th cervical. E, E'. Transverse process, 6th cervi- cal. F. Embedded arrowhead. F'. Tip of arrowhead. 138 Annals of Carnegie Museum vol. 35 Hibbard, Claude W. 1955. The Jinglebob interglacial (Sangamon?) fauna from Kansas and its climatic significance. Contributions from the Museum of Paleontology, University of Michigan, Ann Arbor, v. 12, no. 10, p. 179-228. Parker, John D. 1953. Bone caves in Pennsylvania. The American Caver, December, p. 10-14. (National Speleological Society, Bulletin 15.) Rhoads, Samuel N. 1903. The mammals of Pennsylvania and New Jersey. Privately print- ed, Philadelphia. 266 p. Roslund, Harry R. 1951. Mammal survey of north central Pennsylvania. Final Report Pittman-Robertson Project 37-R. 55 p. Pennsylvania Game Commission, Harrisburg, Pa. Stone, Ralph W. 1932. Pennsylvania caves. Pennsylvania Geological Survey, ser. 4, Bul- letin G3, ed. 2. p. 1-135. 1953. Descriptions of Pennsylvania’s undeveloped caves. The Ameri- can Caver, December, p. 51-137. (National Speleological So- ciety, Bulletin 15.) .P V JVTHSO^^ ART. 10. ARCHAIC HUNTERS UPPER OHIO VALLEY* By Don W. Dragoo 'v Assistant Curator of Section of Man, Carm PREFACE The Archaic cultures of the eastern United States are of great importance in the understanding of the peopling and cultural development of the New World. Until recently, most of our knowledge of the Archaic was based on the work of William S. Webb, T. M. N. Lewis, and Madeline Kneberg in Kentucky and the Tennessee Valley, and the studies of William Ritchie in New York state. In the vast region between these two distant areas little was known of the Archaic cultures. The cultures described for the two known areas appeared to be so different that only vague references were made to their similarities. This report presents the results of recent studies of Archaic cultures con- ducted by Carnegie Museum, Pittsburgh, in the Upper Ohio Valley. This region, encompassing portions of the states of Pennsylvania, Ohio, New York, West Virginia and Maryland, is a key link between Kentucky and Tennessee Valley Archaic and Northeast Archaic. Answers to the following questions were a prime purpose of this project. 1. What anthropological concepts are important to the understanding and definition of the Archaic? 2. What Archaic cultures were present in the Upper Ohio Valley and what were their cultural, temporal, and spatial relationships to each other? 3. What were the relationships of the Upper Ohio Valley Archaic cul- tures with those of surrounding areas? 4. What were the relationships of the Archaic cultures with the earlier Paleo-Indian? 5. What were the origins and routes of migration for the Archaic cultures into the Upper Ohio Valley and the Northeast? Since most of the Archaic sites in the Upper Ohio Valley are small and contain no depth of debris, the excavation of the Dixon and Rohr shelters produced a sequence of Archaic and later materials which has become an important aid in the chronological ordering of many small sites known only by surface collections. The information gained from the Dixon and Rohr excavations has made it possible to revaluate several recently reported sites. In a number of cases, new interpretations of the cultural and temporal relationships of these sites are stated in this article. Since I began this project in 1954, I have had the valuable aid and assist- ance of many people. I wish to express my appreciation to Albert Bauer, Betty Bauer, John Leppla, and Dorothy Dragoo for their labors in the Dixon and Rohr excavations. William S. Webb, of the University of Kentucky, kindly made available to me artifacts and depth locations of these artifacts for Kentucky Archaic sites. Through the courtesy of Glenn A. Black, of * Portions of this article appeared in “Archaic Cultures of the Upper Ohio Valley,” a doctoral dissertation presented to Indiana University, June 1957. 139 Submitted for publication, January 17, 1958 Issued January 5, 1959 SMITH A BD | "ft 140 Annals of Carnegie Museum vol. 35 the Indiana Historical Society, the materials of the McCain site were loaned to me for restudy. I wish to express my appreciation also to Paul Gebhard, Dorothy Cross, Irving Rouse, William A. Ritchie, William S. Webb and John Witthoft for information on specific problems encountered during the writing of this report. I especially want to extend my thanks to James L. Swauger and Mrs. Dorothy Dragoo for their aid in the preparation of the manuscript. Their criticisms and suggestions were invaluable. INTRODUCTION What is the Archaic? The term “The Archaic” was first used by William A. Ritchie (1932) in his classification of certain early manifestations in New York state. Since its inception the term has seen considerable usage. It has been both accepted and rejected. It is a term that has been subjected to several conflicting inter- pretations. Ritchie’s (1944, p. 321) first extensive formulation of the Archaic was as follows: “The long postulated archaic level in New York recently confirmed by intensive work in the Southeast consists of an aggregate of discrete foci, sharing a hunting- fishing-gathering economy. Its chief characteristics are the absence of horticultural traces, ceramics, and the smoking pipe.” Ritchie (1944, p. 319) presented certain specific elements which he believed typical of the Archaic in New York. “Analysis of the archaic level as a whole shows (a) a large variety and numerical abundance of chipped stone types; (b) the lack of all the so-called “problematical” group of polished stone artifacts, except the bannerstone of several simple forms; (c) a considerable typological range in and a large number of bone tools; (d) the prevalence of copper tools and the total lack of copper ornaments; (e) the general absence of shell artifacts; (f) the complete dearth of pipes; (g) the want of pottery except perhaps in the closing phases; (h) the non-existence of agricultural traces; and (i) the large variety of burial practices, generally not involving mortuary offerings.” Ritchie (1955, p. 3) defined his most recent concept of the Archaic in the Northeast as follows: “I conceive of the Archaic as essentially a cultural stage of several thousand years duration prior to the introduction of ceramics, the smoking pipe and horticultural practices, correlated with a mobile or seminomadic way of life, with small band organization, simple social structure, flimsy settlements, and a food-seeking sub- sistence pattern. I do not suppose that this was abruptly terminated or replaced, or that it everywhere yielded to identical changes, certainly not simultaneously. It is amply evident that the Archaic pattern of existence persisted in certain relatively isolated districts of the Northeast long subsequent to the establishment of the Woodland pattern elsewhere in the area, also that many later manifestations are firmly rooted in this older soil.” Thus for Ritchie (1955, p. 4) the Archaic was essentially a developmental stage between the simpler nomadic Paleo-Indians and the more complicated semi-sedentary way of life developed by the agricultural Woodland groups. Ritchie believes that the Archaic in the Northeast was forest adapted at all times, but it also probably underwent some modifications in response to 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 141 climatic and physiographic changes. For Ritchie the Archaic is not uniform but comprises a complexity and variety in the details of its equipment be- cause of the interaction between dissimilar traditions (Lamoka and Lauren- tian) in the Northeast and with other connections outside the area. While viewing the Archaic as a stage in man’s development in the North- east, Ritchie (1955, p. 9) also called upon local environmental conditions and independent invention as factors in the differentiation of the Archaic into foci or phases. He does not conceive of several regional variations as developing in a vacuum, but stresses the interplay of the process of selective diffusion among semi-isolated and localized centers. Ritchie’s term “Archaic” was readily taken over by several workers in the lower Ohio Valley and the Southeast. After the excavation of many non- pottery sites in the Tennessee Valley, Webb and Dejarnette (1942, p. 319) stated: “Since the first report of a non-pottery site in New York, other such sites have been investigated . . . Ritchie has designated this pattern as Archaic, partly on a basis of stratigraphy and because of the absence in the complex of agriculture and pottery . . . Because it now appears that this non-agricultural, non-pottery, hunter- fisher-collector pattern of culture may have been widespread in Eastern United States in early aboriginal times, the term “archaic” is here adopted to designate this pattern manifestation in Kentucky and Alabama.” Implicit in Webb’s writings is his belief that the Archaic has three basic criteria— age, hunter-fisher-collector economy, and the absence of ceramics. In a report by Webb and Haag (1940, p. 109) it was also implied that the Archaic constituted a level or stage in culture. Other workers in the Southeast such as Lewis and Kneberg (1947, 1955), and Fairbanks (1942) accepted the term Archaic for cultural manifestations representative of a hunter-fisher-collector economy devoid of ceramics. Im- plied in the writings of these workers as well as in the works of Webb is the concept of a culture level or stage with rather definite time or age con- notations. After having excavated many deeply stratified sites in the South- east, it would have been difficult for these workers not to have had some definite concept of age for the Archaic in their formulations. The first major objection to the use of the term Archaic was made by Griffin (1946, p. 42). He believed that the artifact assemblage of Archaic was not typologically Archaic and that a great majority of the pre-ceramic units of the eastern United States were relatively late in the pre-ceramic cultural development of the country as a whole. For Griffin the term Archaic was as much a misnomer for the eastern United States as it was in middle America where it had been inappropriately applied to a late cultural assemblage. The term Archaic should have been reserved for the oldest manifestations in North America when such materials became known. Griffin (1946) in his paper on change and continuity in eastern United States preferred to discuss what other workers were calling Archaic as part of the Paleo-Indian problem. Griffin (1952, p. 352-364) later dropped his opposition to the term Archaic and presented a series of prehistoric periods for the eastern United States which he called the (1) Paleo-Indian, (2) Archaic, (3) Early Woodland, (4) 142 Annals of Carnegie Museum vol. 35 Middle Woodland, and (5) Mississippi and Late Woodland. Under this scheme Griffin also recognized both early and late Archaic periods, each with a number of distinctive subdivisions. For Griffin each of the above men- tioned periods represents a stage or level in the progressive evolution of culture. William H. Sears (1948, p. 122-124) expressed a number of adverse criti- cisms of the use of the term Archaic. His argument was that Ritchie’s de- tailed trait list for the Archaic did not serve to distinguish the complex from Early Woodland levels, and thus it could not be used to establish a distinct Archaic pattern. He believed that the concept of the Archaic as a complex which is non-ceramic, non-horticultural, old, and has a hunter- fisher-collector economy to be “entirely too broad to be of any use.” Sears (1948, p. 124) was viewing the problem of Archaic cultures from the standpoint of its inclusion within the framework of the Midwestern Taxo- nomic Classification established by McKern and other workers in the eastern United States. Important to the function of this classification is the prepara- tion and comparison of trait lists for many sites. Sears prepared trait lists for many sites classified as Archaic which revealed that their relationship to each other was sufficient to include them all in a pattern. He then com- pared his Archaic trait lists with the statement of the First Conference on the Woodland Pattern (Angnymous, 1943, p. 393-400). His conclusion was that there were some variations, but except for ceramics in Woodland there were no differences of importance. In regard to ceramics, Sears pointed out that in many instances Woodland ceramics appear in the closing phases of a manifestation with no accompanying change in the non-ceramic inventory. As examples he gave Ritchie’s Brewerton sites and Webb’s Pickwick Basin sites. Sears’s (1948, p. 124) conclusion was as follows: “It appears then that the manifestations which have been called the Archaic represent the early phases of the continuum known as Woodland. It is not, however, necessary to classify them in the Woodland Pattern. This depends on the criteria for the establishment of, or inclusion in, a pattern, which is after all a classificatory device, and on the fact that addition of non-ceramic manifestations to a complex whose most useful criterion is ceramics, would confuse things considerably. A compromise might be to group these non-pottery sites under some such term as Pre- or Non-Pottery Woodland. This would keep the groups separated termino- logically and at the same time indicate the relationship. In general, I would be inclined to accept the non-pottery section of the trait list offered by the First Conference on the Woodland Pattern as the most useful set of criteria for the delineation of this non-ceramic pattern or section of the Woodland Pattern.” It is evident that in 1948 Sears viewed the Archaic as only the early or beginning phase of the Woodland Pattern. Woodland was seen as a cul- tural continuum with both non-ceramic and ceramic phases. It is, indeed, true that there are sites on which pottery makes its appearance without greatly changing the non-ceramic inventory, but I view such sites as being on the periphery of areas in which the ceramic Woodland cultures were becoming established. Only certain Woodland traits, such as ceramics, had diffused to these areas where they were incorporated into long-established Archaic complexes; There is ample evidence in nearly all of these areas to indicate that the Archaic way of life soon gave way to new ways of living. 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 143 Equally important as the introduction of ceramics, agriculture, smoking pipes, and other new cultural items was the arrival of peoples of a physical type different from that of the Archaic (Neumann, 1952, p. 13-34). Un- doubtedly, many of the peoples and traits of the Archaic became part of this new way of life. In view of the evidence for the introduction of several new and important culture elements as well as new peoples, I certainly hesitate to call the Archaic the non-pottery section of the Woodland Pattern. A recent publication by Sears (1954, p. 23-36) indicated that he also has accepted the Archaic as an established cultural entity. Just as there has been confusion of Archaic and Woodland, there has been confusion of Archaic with Paleo-Indian. The term Paleo-Indian was origi- nally used by Frank H. H. Roberts (1940, p. 51-116) for American cultural assemblages which appeared to be chronologically early on the basis of geologic, faunal, or typological evidence. As was previously mentioned, Griffin (1946, p. 42) extended Roberts’s use of Paleo-Indian to include com- plexes generally known as Archaic. John Witthoft (1952, p. 465) took note of this confusion and stated the following: “The term Archaic may lead to some confusion because this word has been used to include what I call Paleo-Indian, even as Paleo-Indian has been used to include what 1 call Archaic. I think of the local archeology as classified first in a chrono- logical framework and only after that in terms of intra-areal relationship. Such an outline includes three major technological stages, which equate, to the best of our knowledge, with major time divisions. The Paleo-Indian Epoch, so far not divisible into periods, lacked stone pecking and grinding techniques. The Archaic Epoch, provisionally divided into two periods, has as its ‘index fossils’ ground stone tools but lacks pottery. The Woodland Epoch, now seen as four or five periods, was characterized by pottery-making. This triple division seems to be valid almost throughout the continent; it is not to be considered as a major tenet of American archeology, but rather as a tool for segregating cultures into natural classes for further study.” Witthoft finds good technological reasons for separating Archaic from either Paleo-Indian or Woodland cultures. He sees three major technological advances which have chronological significance in the progressive evolution of culture in the New World. Each of these divisions he would term an “Epoch,” but he cautions that these divisions are only tools for segregating cultures into “natural classes” for study. Witthoft (1952, p. 465) rejected the Midwestern Taxonomic System as a method for ordering archeological data. Thus, he did not fall into the statistical pitfall that led Sears (1948) to proclaim Archaic as the early non-ceramic part of a Woodland continuum. A survey of recent literature indicated that the term “Archaic” has sur- vived the many criticisms leveled against its use. To use Witthoft’s term, it has become an accepted “Epoch” in the culture history of the eastern United States. During the past few years many contributions have been made which have aided the understanding of the content and form of many manifestations of the Archaic. Equally important has been the attempt to state and define the theoretical foundations on which the Archaic, or any other culture “Epoch” or “stage,” is based. In this matter, a recent paper by David A. Baerreis (1955, p. 1) warrants attention. Baerreis gave his definition of an Archaic “stage” as being represented by 144 Annals of Carnegie Museum vol. 35 cultures with a hunting and gathering economy and lacking horticulture. The appearance of ceramics in an area was taken as a terminal date or horizon marker as cultures representing an Archaic stage may persist even into historic times. Baerreis believed that the appearance of horticulture constitutes an alternative and more logical terminal horizon marker than the appearance of pottery but he pointed out that the initial acceptance of agriculture is difficult to recognize in the present state of archeological knowledge. Baerreis also called attention to the fact that the term Archaic can be used in at least two primary senses. First, it may have reference to a specific stage of cultural development characterizing a regional sequence. Second, Archaic may be used in reference to a cultural tradition which represents a particular configuration of culture traits that persist through a number of culture stages. Baerreis warned against speaking of an archaic tradition for he believed that the archaic stage may comprise many recognizable cultural traditions. The spread of distinct traditions or the diffusion of specific ele- ments from these traditions may have created the impression of develop- mental sub-stages within the archaic. It is evident that Baerreis took as granted the existence of culture evolu- tion. His acceptance and use of the term stage indicated his belief in the progressive development of culture through time. He was, however, very much aware of factors such as diffusion and environmental influences in the shaping of any culture of a specific area. Speaking of the differences seen in the Archaic Baerreis (1955, p. 10) stated: “No single answer can be given if a question as to why these differences appear is raised. A simple solution might be to say that we are dealing with a later time period when a higher level of cultural growth and complexity had been attained. Yet radiocarbon dates are pushing the limits of Archaic cultures farther back in time so that at present time we are beginning to see a clear temporal overlap in the two broad culture types (Paleo-Indian and Archaic). Differences in environ- mental factors may also be posed and certainly they do have their influence on the total configuration. Yet it would be pushing an already weak explanation to the point of absurdity to postulate, for example, that the greater diversity of game in the eastern United States could be responsible for a more heterogeneous assem- blage of projectile point types and to extend this reasoning to other aspects of the culture. The problem would appear to largely revolve around differential situations that facilitated acculturation in one region in contrast to the other.” As to how differential situations have aided or retarded acculturation Baerreis (1955, p. 10) suggested that the differences that set apart Paleo- Indian and Archaic culture types are perhaps due to basic differences in social structure. The economy of Paleo-Indian groups was based upon hunt- ing. It is assumed that there were a number of widely scattered patrilocal bands. Because only a small number of male adults would have been manu- facturing projectile points the techniques would tend to have been perpetu- ated within the patrilocal band. Only women would have been directly in- volved in bringing in new ideas from the outside. Among Archaic groups, seed gathering was of importance as well as hunting. The importance of women, who are usually the gatherers, in the economy may have caused a shift from patrilocal to matrilocal, or at least bilocal, bands. The marriage 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 145 pattern would then have brought in new males to the local group and with them new ideas concerning styles of projectile points and other items. Griffin (1952, p. 354) also attempted to reconstruct the social organization of the Archaic; but, unlike Baerreis who saw women’s part as collectors as highly important in the Archaic, he postulated a patrilineal descent for the Archaic on the basis of the male dominance in hunting and fishing activities. I look favorably upon Baerreis’s suggestion of the presence of matrilocal or at least bilocal bands in the Archaic. Most Archaic sites produce abundant evidence (charred seeds and nuts, grinding stones, etc.) pointing to the importance of seed gathering and other collecting activities. Before passing on to a working definition of the Archaic as a guide to this present study, it is necessary that we look more closely at the concepts of “stage,” “tradition,” and “co-tradition.” As has been noted in previous paragraphs, these concepts are being strongly emphasized by present-day American archeologists and are intimately involved in any definition of the Archaic. The idea of a series of stages as reflecting the chronological development of culture is not new to archeology. Following the work of Morgan (1877) and other evolutionary ethnologists of the nineteenth century, the idea of stages became exemplified by the Stone-Bronze-Iron classification used by students of European prehistory. The facts of cultural distribution as now known have led largely to the abandonment of evolutionary schemes as a means of presenting chronology, but such schemes are still found useful in expressing the degree of cultural development. In the New World the evolutionary approach to chronology was recently used by Strong (in Bennett, 1948, p. 98) in the Viru Valley of Peru where he used a series of “cultural epochs” or stages of cultural development which he believes also applicable to many other areas of the world. Rouse (1953, p. 72) pointed out that Strong’s formulation has two advantages over previous studies of the same kind: “ (1) in so far as possible, it is based upon observed distributions rather than on conjecture, and (2) for the most part it is defined in terms of degree of develop- ment rather than of specific cultural content. It therefore provides a means of expressing the broad sweep of chronology over a whole series of areas without bringing in the differences in detail from area to area.” This view of the cultural stage as outlined above by Rouse is like the one expressed by Baerreis in his formulation of an Archaic stage. Griffin (1952) and his associates took a somewhat similar position in their book “Archeology of Eastern United States.” Another concept currently receiving wide use by American archeologists is that of the “tradition.” Kroeber (1944, p. 108-111) applied the term “horizon style” to the phenomenon of one or more traits extending from area to area through a series of contemporaneous periods. Willey (1945) applied the term “tradition” when one or more culture traits survived from period to period within a single geographic area. The “tradition” for Willey was the counterpart of Kroeber’s “horizon style.” Willey (1953, p. 374) sum- marized his concept of “tradition” as follows: “The theory underlying the tradition concept is well expressed in the term 146 Annals of Carnegie Museum vol. 35 itself. For “tradition” implies deep-set and channeled activity or patterned ways in which the vitality of a culture expresses itself in strong preference to other possible ways. The conditions surrounding this rigidity of expression which results in the long-time traditional expression are an interesting problem for future in- vestigation. In dealing with such things as pottery decoration, the archeologist is undoubtedly investigating what is relatively trivial in past human events. The failure of polychrome painting to take hold on the northern coast of Peru in the face of the white-on-red tradition, despite several attempts to introduce it, may eventually be revealed as nothing more mysterious than an absence of suitable mineral pigments in that part of the country. On the other hand, the unraveling of seemingly insignificant threads in an attempt to factor out causality may lead us to a greater understanding of ‘tradition set’ and ‘tradition persistence’ in institutions which loom larger in human affairs.” Related to the concept of tradition is the idea of the area co-tradition (Bennett, 1948). The co-tradition is based upon the persistence of a number of closely interrelated traditions found within a specified area. It presents a means by which the archeologist can plot the history of cultural continuity within the confines of an area. As Willey (1953, p. 374) pointed out, the co- tradition is the culture area with time depth. Although the concepts of “stage,” “horizon style,” “tradition,” and “co- tradition” were formulated by workers in Andean archeology, these terms are being applied by workers in North America. Martin and Rinaldo (1951) used the co-tradition concept with success in the southwestern United States, and Willey (1953) conceived of three co-traditions consisting of the Archaic, Woodland, and Mississippian in eastern North America. In this study of the Archaic in the Upper Ohio Valley and its relationship to other cultural manifestations of eastern North America, I have been guided in the selection of materials and sites by the following points of definition for the Archaic: 1. The Archaic is represented by cultures with a hunting, fishing, and gathering economy, lacking agriculture. 2. The appearance of ceramics in the area is taken as a terminal date for the Archaic. It is recognized, however, that many Archaic traits may per- sist through a number of cultures which possess ceramics and agriculture. 3. The lower time limit for the Archaic is difficult to define clearly because the final phases of Paleo-Indian cultures were often involved in acculturational situations with early Archaic cultures. In agreement with Baerreis (1955, p. 2), it seems “preferable to leave the lower limits for the Archaic flexible and to simply attempt to trace the cultures of the delimited stage as far back in time as possible.” 4. The Archaic is conceived as a “stage” or “level” in the development of culture in eastern North America. For practical taxonomic purposes the Archaic can be studied as if it were independent of lower or higher levels, but it is imperative that interrelationships between the levels should not be forgotten. The Archaic is not completely intelligible apart from reference to the levels below and above it. A higher level, however, can not be reduced to a lower level merely because it shares traits with the latter. 5. The Archaic is not a “tradition” but a stage in cultural development in which a number of “traditions” may be recognized. Any tradition or par- 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 147 ticular configuration of culture traits may persist through a number of culture stages. 6. Differences in environmental factors and acculturational situations are important considerations in the understanding of any particular Archaic manifestation or group of manifestations. A major purpose of this report is an attempt to discover and define the processes and factors involved in what we know as the “Archaic.” In the following sections a number of recently discovered Archaic manifestations in the Upper Ohio Valley are described and compared with others known in eastern North America. The theoretical concepts just discussed present a framework upon which an understanding of the Archaic can be constructed. THE DIXON ROCK SHELTER While I was on a Carnegie Museum survey trip to the Cheat River region of northeastern West Virginia in early November 1953, an interesting rock shelter on the Preston County farm of Jesse Dixon was brought to my atten- tion by Albert H. Bauer of McKeesport, Pa. This site was named for the owner, Dixon, and given the Upper Ohio Valley Archeological Survey num- ber of 46Pr6. The shelter protected a considerable area from rain and chill- ing winds. On the floor was the scattered debris of both Indian and white occupation. According to local residents the shelter was once used as a base by horse thieves, and only a few years ago a destitute family spent a winter under the cliffs protection. The geographic location of the shelter and the materials found on the surface of its floor were of special interest. The site being near the limits of the Ohio River drainage and in close proximity to the upper reaches of the Potomac Valley presented this problem. Would the cultural materials be typical of those found in the Upper Ohio Valley or would they resemble more closely those of the Potomac area? Here was an opportunity to observe cultural movements at the Ohio Valley’s rear door. An examination of the few artifacts gathered from the shelter gave evidence of cultural change among the types of pottery and projectile points. Typologically there were objects ranging from the Late Prehistoric back to non-ceramic times. The dark color and apparent depth of the soil in the shelter coupled with the varied artifact types gave us hope that we had found a good stratified de- posit. If this were the case, we would have a good check on the sequence and development of human habitation on the eastern fringe of the Ohio Valley. With the kind co-operation of the owner, arrangements were made for Carnegie Museum to excavate at the shelter during the last week of July and the first two weeks of August, 1954. Mr. Bauer’s tireless participation in the excavations and the processing of recovered materials did much to make the project a success. This report gives the results of our partial ex- cavation of the site. NATURAL SETTING The Dixon Shelter is located in mountainous country at an elevation of 1800 feet above sea-level. Many hilltops in the surrounding area reach an 148 Annals of Carnegie Museum vol. 35 elevation of from 2500 to 3000 feet. The overhanging ledge or roof of the shelter is composed of the Pottsville conglomerate, the basal formation of the Pennsylvania period in the area (Hennen and Reger, 1915, p. 177-184). It is a compact, grayish-white sandstone with many quartz pebbles. Under- lying the conglomerate is a fine brown sandstone which has eroded extensive- ly to form the two chambers of the shelter. Approximately one hundred feet in front of the shelter and responsible for its formation is a small, fast- flowing stream known as Lick Run. This stream has moved east and cut its new channel about four feet into bedrock since it eroded the brown sand- stone. About one-fourth mile south of the shelter, Lick Run enters Roaring Creek which empties into the Cheat River 33A miles to the southwest. The Cheat River drains a large portion of eastern West Virginia into the Ohio River system. From its source at an elevation of 4000 feet to its mouth where it is only 780 feet, the Cheat is a turbulent mountain stream with many rapids. The area is covered with fine forests of hardwoods. Only a small portion of the land is suitable for farming and has been cleared. On the hilltops and slopes are great stands of white, black, scarlet, and chestnut oaks mixed with red maple, elm, and sweet birch. Along the small streams in the valleys are stands of hemlock crowded by thick growths of mountain laurel and rhododendron. Sassafras, silky dogwood, black cherry, and sour gum trees are scattered throughout the area. There are many shrubs and bushes with edible nuts, fruits, or berries. Among these are the hazelnut, elderberry, huckleberry, dewberry, blackberry, partridge-berry, deerberry, mountain tea, and the late low blueberry. Plants of the wild hydrangea and sunflower and of Indian tobacco (Lobelia inflata ) were seen growing near the shelter. In prehistoric times the Cheat River region abounded in animal life. Deer, elk, bear, bob-cat, mountain lion, rabbit, squirrel, raccoon, and many other species were common. In the streams were black bass, sunfish, brook trout, catfish, suckers, and snapping turtles. Fresh-water clams were scarce as the streams are too swift and stony. Wild turkey and grouse were common and some waterfowl may have inhabited the marshy ponds cradled among the hills. EXCAVATION The overhanging roof of the Dixon Shelter extends approximately 150 feet along the outcrops in a line from southwest to northeast. From the front of the shelter to the rear wall, the recess varies in depth from a few feet to as much as 25 feet. Near the front of the shelter the ceiling ranges in height from 6 to 10 feet above the floor, but it slopes gently to a height of only two to four feet at the rear wall. Near the center of the shelter is a large fissure in the conglomerate stratum and a resulting breakdown of the ceiling. When this breakdown occurred is not known but it is probable that it has been relatively recent since cultural materials were found under the edges of the fallen boulders. The collapsed ceiling and the boulders strewn on the floor divide the shelter into two natural divisions at the present time. Our excavations were confined to areas not covered by the large boulders since we had no convenient method of removing them. (See Fig. 1.) 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 149 The area of the shelter was surveyed, and a base-line 150 feet in length was established parallel with the front of the shelter. This line was then divided into three divisions each 50 feet in length with division A at the northeast end of the shelter and division C at the southwest end. Division B was at the center and almost entirely covered by the large boulders. A grid system consisting of five- by five-foot squares was established in each division. The horizontal location of all materials removed in the excavations was recorded by division and block designation as well as vertical level. (See Fig. 2.) A datum point was established and vertical measurements were made in reference to this point. Each excavated block was dug by arbitrary six-inch levels. The levels of all the blocks dug in division A were designated in relationship to the datum point. When the soil profiles and the recovered artifacts were observed it was apparent that the relative position of any artifact-bearing level with the datum point was of little significance. The main vertical reference for any object or level of objects was its relative dis- tance from the surface and the depth of the deposit for the particular block in which it was found. For example, the debris deposits in the blocks of division A, near the center of the shelter were as much as twice as deep as the deposits in the blocks near the northeastern end, but the relative position of cultural materials remained the same. With this information, all levels of division A were converted into straight sequence from the measure- ments made from the datum point, and all levels excavated in divisions B and C were determined in sequence from the surface. This greatly simplified Fig. 1. General view of Dixon Shelter before excavation 150 Annals of Carnegie Museum vol. 35 the handling of the materials when the detailed study was begun. All excavating was done by customary archeological methods. Trowels were used for nearly all primary digging. Many objects were recovered as the soil was loosened but all earth was passed through a screen as a safeguard since available light did not permit proper viewing. Observations of soil differences, features, and cultural materials were recorded for each level within a block and, where necessary, drawings and photographs were made. All open excavations were refilled when work was completed. Fig. 2. Dixon Shelter. Division A excavations in progress Each major area excavated presented its own peculiar problems. In the following paragraphs the excavation in each division is discussed in detail. Division A Division A included all of the clear, excavatable area of the shelter north- east of the collapsed ceiling. Digging began near the margin of the fallen boulders and continued to the northeast end of the shelter, a distance of approximately forty feet. Immediately below the edge of the overhang and on the floor of the shelter was a deposit of material broken from the ceiling and washed from the surface above the outcrop. This deposit appeared as a ridge following the outline of the overhang. Testing showed this area deposit to be stony and relatively clear of cultural debris. Major digging was confined to the floor area between this ridge of rubble and the rear wall of the shelter. Ten whole or partial blocks were dug in division A. The cultural debris 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 151 in the blocks nearest the center of the shelter adjoining the collapsed ceiling reached a maximum depth of 34 inches below the surface. Underneath the debris-laden, dark-gray soil, was found a fine yellow gravel. No cultural materials were found at any depth in this gravel but some objects were definitely resting on its surface. The yellow gravel represents the floor of the old stream bed at the time erosion was cutting away the soft sandstone outcrop. Since there was no sterile layer of humus-laden soil between the yellow gravel and the cultural deposits, the shelter was probably occupied by people soon after the channel of the stream moved away from the stone outcrop. The debris in the blocks extending towards the northeastern end of the shelter became progressively thinner in depth until in the last excavated block it was at a maximum of 18 inches. In this area the yellow gravel was thin and rested upon the sandstone bedrock which was closer to the surface in this area because the formation dipped gradually downward to the southwest. All of the dark-gray soil of division A contained cultural debris. This debris varied in kind and age from the surface to the bottom regardless of the total depth. Tools of flint were common throughout this debris but pottery was found only in the upper levels. In blocks with a deep deposit, pottery was found to a depth of approximately eighteen inches below the surface, but in the blocks with the thin deposits pottery was confined to the upper six inches. From one-third to one-half of the lower depth of the deposits was non-ceramic in content. The types of pottery and projectile points found in the various levels and the cultural time periods to which they belong will be discussed in later paragraphs dealing with these objects. Suffice it to say here that definite cultural stratigraphy did exist. No clear- cut soil differences could be correlated with the cultural changes. The debris appears as a gradual build-up with no disturbances such as flood deposits to interrupt the continuity. All the features found in division A were poorly defined. The dark, dry, loose soil does not cut along sharp lines but crumbles and breaks. Thus, a pit dug into the soil did not maintain a sharp outline, and unless the yellow gravel was reached there was little or no change in soil coloration. All of the seven refuse pits recorded were small and shallow in depth. The largest was 18 inches in diameter and the maximum depth was 12 inches. Refuse deposited in these small pits consisted mainly of broken animal bones, flint chips, and pottery sherds. There were four areas in division A where fires of considerable intensity had been located. None of these areas could be called fire-pits since the fires had not been confined to a depression in the soil but had been placed directly on the surface. Around these areas ranging in diameter from 18 to 30 inches was an accumulation of ash, small bits of charcoal, and reddened earth. Small fires of short duration probably were represented by the scat- tered ash and flakes of charcoal found throughout the debris in the shelter. The remains of small fires would have easily become scattered in the loose soil. 152 Annals of Carnegie Museum vol. 35 Divisions B and C The work done in division B will be considered in the overall picture gained from the excavation in division C. Only one block adjoining division C was dug in division B because most of the area was covered by large boulders. Five whole blocks and a portion of a sixth were dug in division C. All of the blocks were well under the protection of the overhang towards the southwest end of the shelter. One large slab had collapsed from the ceiling and covered a portion of the area. This fall appeared to be recent since cultural materials and dark, debris-laden soil were found well under the edge of the stone. The dark, gray soil containing cultural debris extended to a depth of 30 to 34 inches below the surface in all the blocks of this area. Directly below this debris was the yellow gravel as encountered in division A. There was no sterile soil between the yellow gravel and the debris-laden soil. Some objects were found directly on the gravel level but none was found below it. The sequence of cultural materials was confused in a portion of this area because of the presence of a large refuse pit approximately six feet in diameter. This pit was made late in the occupation of the shelter as it began near the surface and extended to a depth of nearly 28 inches. Its contents consisted mainly of elk and deer bones mixed with pottery sherds repre- sentative of various time levels. The margins of this pit, like those found in division A, were indistinct and difficult to follow because of the loose soil. There was the chance that some objects belonging to the pit would be in- cluded in the undisturbed levels surrounding it. All pottery was confined to the first 18 inches of the deposit except in the disturbed area of the large pit where a few sherds were found to a depth of 26 inches. Flint objects were found at all levels from the surface to the yellow gravel. The large pit was the only well defined feature found in divisions B and C. Some thin lenses of ash were encountered but there was no evidence of the rock-lined fire-pits common to open village sites. POTTERY Pottery was not abundant in the refuse of the Dixon Shelter but a suf- ficient number of sherds was found to give a relatively clear picture of the time sequence and types involved. From all the excavated blocks 681 sherds were recovered. No whole or completely restorable vessels were found. All sherds from the levels of each block were sorted and classified according to “wares” based on the kind of temper. Using Carnegie Museum’s Upper Ohio Valley Archeological Survey’s pottery nomenclature (Mayer-Oakes, 1955a), I identified four major “wares”— shell-tempered Monongahela, lime- stone-tempered Watson, medium grit-tempered Mahoning, and coarse grit- tempered Half-Moon. In the case of the Half-Moon ware, thickness and surface treatment were important factors in separating it from the Mahoning ware, also grit-tempered. Each of the ware groups was divided into its component types with the following results: Monongahela Cordmarked Method of manufacture. Coiled; malleated with a cordwrapped paddle. 1958 Dr a goo: Archaic Hunters of the Upper Ohio Valley 153 Temper. Moderate to finely crushed particles of shell composing approxi- mately 25% of the paste. Texture. Finely laminated. Hardness. 2.5 to 3.5. Color. Exterior surfaces range from buff to black. Interiors usually are gray to black. Surface finish. Exterior surfaces covered by the impressions of a cord- wrapped paddle. These impressions usually meet the lip at an oblique angle. Some sherds show smoothing of the cord impressions. Interior surfaces are smooth. Decoration. Confined to the lip, small indentations of a cordwrapped stick or paddle edge are common. Form. Globular jars or bowls with flattened or rounded lips on a straight or slightly outflaring rim. No appendages noted. Thickness. 4 to 7 mm. Diagnostic features. Paste, surface finish, rim form. Geographic range .. This type is known from all major streams of the Upper Ohio Valley. It occurs most abundantly on sites along the Monongahela River and down the Ohio River proper where it overlaps with similar types of Fort Ancient pottery. Probable relationship. This pottery type is the most common and basic shell-tempered type found in the Upper Ohio Valley. It occurs at villages of the Late Prehistoric Period which in the Upper Ohio Valley appear to be mixtures of Mississippi and Woodland cultures. This is known as the Monongahela Complex and its closest relationship is with Fort Ancient. Size of sample. 102 sherds. References. Butler (1939). Dragoo (1955). Mayer-Oakes (1955). Watson Cordmarked Method of manufacture. Coiled; malleated with a cordwrapped paddle. Temper. Crushed fragments of limestone making up from 25 to 50% of the paste. Texture. Compact but often irregular because of the large size and the angular shape of the limestone tempering. Hardness. 2.5 to 3.5. Color. Both surfaces usually a dull gray or tan. Surface finish. Exterior surface impressed with the marks of a cord- wrapped stick running at an oblique or right angle to the rim. The in- teriors are generally smooth but faint cord impressions occur on a few sherds. Decoration. Confined to the lip or neck area were impressions of a cord- wrapped stick or paddle edge applied at an oblique angle. Form. Globular vessels with straight to slightly flaring rims. Thickness. 4 to 8 mm. Diagnostic features. Paste, surface finish, rim form. Geographic range. This type is known at scattered sites throughout the Upper Ohio Valley. It is best known from sites in the Ohio Valley 154 Annals of Carnegie Museum vol. 35 proper where the type site, Watson, is located. Probable relationship. This pottery type is the one most commonly asso- ciated with Middle Woodland occupations in the Upper Ohio Valley. At the type site it is associated with a stone mound burial complex lying stratigraphically above an older Early Woodland occupation. It is thus representative of the Upper Ohio Valley’s diluted Hopewellian manifestations. Watson Cordmarked overlaps and shares many features with the older Half-Moon type. It remained in use up until Late Pre- historic times when it was replaced by the shell-tempered Monongahela types. Size of sample. 483 sherds. References. Dragoo (1954, 1956). Mayer-Oakes (1955a). Watson Plain This type is similar to Watson Cordmarked in all known features ex- cept for the smooth surface finish. It occurs as a minority companion to the cordmarked type at all sites where it has been found. Its dis- tribution and relationships parallel those of the cordmarked type. Size of sample. 13 sherds. References. Dragoo (1954, 1956). Mayer-Oakes (1955a). Watson Incised This type is not common. It is an incised decorated version of either the Watson Cordmarked or Watson Plain types. The only two sherds of this type at the Dixon Shelter had incised lines over smoothed cord- marking. References.. Dragoo (1954, 1956). Mayer-Oakes (1955a). Mahoning Cordmarked Method of manufacture. Coiled; malleated with a cordwrapped paddle. Temper. Crushed fragments of rock, usually of igneous origin, and some- times chert and quartz. Temper fragments are small to moderate in size and make up 25 to 40% of the paste. Texture. Compact and regular except for a few sherds containing large temper fragments. Hardness. 2.5 to 3.5. Color. Ranges from dark gray to tan. Surface finish. Exterior surface covered by the impressions of a cord- wrapped paddle applied vertically to the rim. The interiors are smooth except for a few sherds which have faint cord impressions. Decoration. Confined to the lip and consisting of the impressions of a cordwrapped paddle. Form. Probably globular with a slightly flaring rim with square to rounded lips. Thickness. 4 to 7 mm. Diagnostic features. Paste, temper size, and surface finish. Geographic range. This type is known from sites throughout the Upper Ohio Valley. It is concentrated on sites in the Beaver River and Ma- honing River valleys but it occurs as a minority type on most sites indicating extensive and long occupation. 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 155 Probable relationship. This type covers a wide range of time and cul- tural differences. It is known stratigraphically to be contemporary with late Early Woodland types and it continued to be made until Late Prehistoric times. As a time marker it is almost useless. It is the Upper Ohio Valley’s contribution to the amorphous type often called “Wood- land Cordmarked”. Size of sample. 25 sherds. References., Dragoo (1954, 1956). Mayer-Oakes (1955a). Mahoning Plain This type possesses, with the exception of the smooth surface, the same general features, geographic range, and relationships as Mahoning Cordmarked. Size of sample. 15 sherds. References. Dragoo (1956). Mayer-Oakes (1955a). Half -Moon Cordmarked Method of manufacture. Coiled; malleated with a cordwrapped pacldle. Te?nper. Coarsely crushed fragments of igneous rock most common, but chert, quartz, and limestone also used. Sometimes a combination of tempering particles used. Temper fragments are large and make up from 50 to 70% of the paste. Texture. Rough and irregular. Hardness. 2.5 to 3.5. Color. From tan to light gray. Surface finish. Exterior surface deeply marked by cord impressions. In- teriors may be smooth or marked by cord impressions. Decoration. No evidence. Form. Large globular vessels indicated. Thickness. 8 to 18 mm. Diagnostic features. Paste, temper size, and thickness. Geographic range. This type is known best from sites on the Ohio River proper but some examples have been found at sites scattered throughout the Upper Ohio Valley drainage area. Probable relationships. This type is the earliest known pottery from the Upper Ohio Valley. When found in stratified sites, it is always at the lowest levels. It has been definitely associated with Adena burial mounds and villages attributed to that culture. Half-Moon Cordmarked is closely related to Baumer (Cole, 1951) and Crab Orchard (Maxwell, 1951) types and the type Fayette Thick (Griffin, 1943) so common in the central and lower Ohio Valley. It also shares basic similarities with Vinette I of New York state (Ritchie and MacNeish, 1949). Size of sample. 37 sherds. References. Cole (1951). Dragoo (1954, 1956). Griffin (1943). Maxwell (1951). Mayer-Oakes (1955a). McKees Rocks Plain This type closely resembles the paste features of the Half-Moon Cord- marked type. The major difference lies in the smooth exterior and in- terior surfaces. In the Upper Ohio Valley this type appears as the com- 156 Annals of Carnegie Museum vol. 35 panion to the cordmarked variety in both time and cultural relation- ships. It is most similar to “Adena Plain” (Griffin, 1942) of the central Ohio Valley. Size of sample. 4 sherds. References. Griffin (1942). Mayer-Oakes (1955). Scarem Plain This type is represented by only two sherds at the Dixon Shelter. Both sherds are untempered and are from small “toy” vessels. Sherds of this type are common on Late Prehistoric sites of the Upper Ohio Valley. Size of sample. 2 sherds. References. Dragoo (1953). Mayer-Oakes (1955a). Pottery of division A. In the blocks of division A, 434 sherds were found. The first level (0-6 inches) contained 267 sherds (61.6%), the second level (6-12 inches) had 134 sherds (30.9%), the third level (12-18 inches) possessed 29 sherds (6.7%), and the fourth level (18-24 inches) accounted for only 4 sherds (0.8%) which were believed possibly to be in the bottom of an ill- defined refuse pit. There were 51 shell-tempered Monongahela Cordmarked sherds of which 84.3% (43 sherds) were found in the first level and the remaining 15.7% (8 sherds) in the second level. All of the shell-tempered sherds recorded for the second level were recovered from the blocks containing the deepest debris. As the blocks of division A became thinner in depth toward the northeast end of the shelter, the shell-tempered sherds were found only on or just below the surface. (See Fig. 3, rows 1-2.) The 347 Watson Cordmarked and Watson Plain sherds found in division A were distributed 61.9% (215 sherds) in the first level, 32% (111 sherds) in the second level, 4.9% (17 sherds) in the third level, and 1.2% (4 sherds) in the fourth level. Thus, 38.1% of the limestone-tempered pottery was found below the first level in contrast to only 15.7% for shell-tempered sherds. (See Fig. 3, rows 3-5.) Of the 22 Mahoning Cordmarked and Mahoning Plain sherds removed from division A, there was 27.2% (6 sherds) in the first level, 36.4% (8 sherds) in the second level, and 36.4% (8 sherds) in the third level. There was 72.8% of the sherds of this type evenly distributed in levels two and three below the first level in comparison with 38.1% of limestone-tempered and only 15.7% of shell-tempered sherds. The meager samples of Half-Moon Cordmarked and McKees Rocks Plain types were distributed 8.3% (1 sherd) to the first level, 58.3% (7 sherds) to the second level, and 33.4% (4 sherds) to the third level. Four of the seven sherds listed for the second level were from blocks where the debris was thin. These sherds were at the bottom of the ceramic zone. (See Fig. 3, rows 6-7.) Pottery of divisions B and C. The contiguous blocks of divisions B and C were considered as a single unit for pottery analysis since the nature and depth of the debris appeared nearly identical in all blocks. The sequence of deposition was interrupted only by the large refuse pit previously mentioned. This large, vaguely defined pit made proper placement of a few sherds 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 157 Fig. 3. Dixon Shelter. Rows 1-2, Shell-tempered pottery. Rows 3-5, Lime- stone-tempered pottery. Rows 6-7, Heavy grit-tempered pottery 158 Annals of Carnegie Museum vol. 35 very difficult since it was almost impossible to draw the line between the disturbed soil and the naturally accumulated soil. The same pottery types found in division A were present in divisions B and C. Of the 248 identifiable sherds found in divisions B and C, there were 175 sherds (70.6%) in the first level, 57 sherds (23.0%) in the second, 10 sherds (4.0%) in the third, and 6 sherds (2.4%) in the fourth level. All the sherds tabulated for the fourth level were found in the immediate area of the large refuse pit and may have been included in that disturbance. There were 52 shell-tempered Monongahela Ware sherds of which 90.4% (47 sherds) were in the first level and the remaining 9.6% (5 sherds) in the second level. Distribution of the 149 limestone-tempered Watson Ware sherds was 69.8% (104 sherds) in the first level, 24.8% (37 sherds) in the second level, 4.0% (6 sherds) in the third level, and 1.4% (2 sherds) in the fourth level. There were only 18 grit-tempered Mahoning Ware sherds of which 33.4% (6 sherds) were in the first level, 22.2% (4 sherds) in the second level, 22.2% (4 sherds) in the third level, and 22.2% (4 sherds) in the fourth level. The 29 coarse grit-tempered Half-Moon Ware sherds were confined to the upper two levels. There was 62.0% (18 sherds) of the sherds in the first level and 38.0% (11 sherds) in the second level. Nearly all of these sherds were found in the area disturbed by the refuse pit and were, thus, out of the natural deposition sequence. POTTERY SUMMARY Pottery types typical of all the major ceramic periods of the Upper Ohio Valley were found in the Dixon Shelter. The pottery sample was small and from a too shallow depth to be of the highest value for stratigraphic pur- poses. Thus, a pottery sequence determined from the sherds at the Dixon Shelter could be only suggestive; but when this material was placed in rela- tion to sequences known from other sites in the Upper Ohio Valley, the pattern was almost identical. At all known stratified sites in the Upper Ohio Valley, the Monongahela pottery types are found only in the top levels and are typical of the Late Prehistoric period. In all areas dug at the Dixon Shelter, Monongahela pottery was confined to the upper two levels with the highest percentage of sherds in the first level near the surface. This pottery represented the terminal Indian occupation. In deep sites Watson Ware pottery occurs stratigraphically below the Monongahela types and is representative of the Middle Woodland period. At the Dixon Shelter, this pottery was the most abundant ware. It occurred from the upper to the lowest pottery-bearing levels. On the basis of type percentages per level, the Watson Ware was in sequence below the Monon- gahela Ware. Its extensive vertical distribution would indicate a relatively long period of its usage at the shelter. Mahoning Ware pottery was distributed rather evenly from top to bot- tom levels. A greater percentage of this type was found in the lower ceramic levels than was the case with Watson Ware. The Mahoning Ware appears as the surviving member of an early grit-tempered ceramic tradition. At 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 159 all known stratified sites in the Upper Ohio Valley, the Mahoning Ware was used during the last phases of the early coarse grit-tempered tradition and continued in use until replaced by shell-tempered Monongahela Ware. It often occurs as a minority ware in association with the Watson Ware (Dragoo, 1956). Half-Moon Ware pottery is the earliest known from stratified sites in the Upper Ohio Valley. It is the typical coarse grit-tempered pottery of the Early Woodland period. The position of the Half-Moon pottery at the Dixon Shelter was not clearly defined. In division A it was found in the lower ceramic levels, but in divisions B and C the best examples of it came from a disturbed area and in the upper levels. The few sherds found in division A seem to indicate that Half-Moon Ware was also the earliest pottery at the Dixon Shelter. CHIPPED-STONE ARTIFACTS The same problems encountered in the study of the pottery were present in the tabulation of the chipped-stone materials. The varying depth of the habitation debris made it impossible to correlate objects from all blocks on a straight level basis. In order to approximate more closely the actual cul- tural stratigraphy for study purposes, the blocks were divided into two groups. All blocks containing four or more levels (most blocks had five levels) were put in one group and all those containing only three levels were placed in another. By doing this, it was possible to compare accurately the sequence of chipped-stone materials for the entire excavation and their relationship to the pottery-bearing levels. Projectile points The most abundant chipped-stone items recovered from the excavation were projectile points. From all the blocks, 187 points were collected. Of these, 180 could be sorted into generally recognized types. The seven re- maining points were of unusual or undeterminable form. Of the 134 points found in the blocks containing four or more levels, the distribution was 27 (20.15%) in the first, 33 (24.62%) in the second, 34 (25.37%) in the third, 33 (24.62%) in the second, 34 (25.37%) in the third, 33 (24.62%) in the fourth, and 8 (5.97%) in the fifth level. In the blocks containing only three levels there were 53 points distributed 12 (22.64%) in the first, 23 (43.39%) in the second, and 18 (33.98%) in the third level. Of the 180 classified points, there were 124 points (69.4%) found in levels which also produced pottery and 55 points (30.6%) in levels where pottery was absent. All of the 180 points were sorted into generally accepted types (stemmed, side-notched, corner-notched, and triangular) and plotted as to level distribution. In the discussion which follows, each type of point and its subdivisions will be described and its position in relation to pottery and non-pottery levels noted. Stemmed points. These points constituted 56.11% (101 points) of the classified sample of 180 points. During the sorting, two major variations of the stem were encountered. The first variation consisted of points with a parallel stem and a rounded or square base. The second and most numerous variation was points with an expanded stem. 160 Annals of Carnegie Museum vol. 35 The 36 parallel-stemmed points made up 35.64% of the total stemmed point sample (101 points) and 20.0% of the total point sample. Of these points, 77.78% (28 points) were found in levels where pottery also occurred, while the remaining 22.22% (8 points) had no pottery association. The stemmed points were the largest projectile points found in the excavation. They range in length from 114 to 314 inches. In Fig. 4, no. 1-10, several of the largest points of this type are shown. All of these points are of a coarse-grained flint except for no. 10 which is of quartz. The chipping on these points is crude, but the nature of the coarse material used in their manufacture would have made fine chipping impossible. In spite of the crude appearance of these points, they were, with the exception of two, found distributed throughout levels containing pottery. The unique shape and the materials from which they were made set them apart from all others and produce speculation for their origin outside of the Ohio Valley. Points of similar type are most common on many sites in the Potomac River drainage. The only points comparable to these in the Ohio Valley occur on the eastern fringe of the area in Garrett County, Md. (Corliss, 1954) where they are commonly associated in an Archaic context more closely related to materials found along the eastern seaboard. A few points in the parallel-stemmed category have rounded bases. These are shown in Fig. 4, no. 12-16. Only the bases of this type of point were found. All are of fine-grained flint and nicely chipped. They are remi- niscent of points found on Early Woodland sites, particularly Adena, in the Upper Ohio Valley. Stemmed points were most common in the lower pottery levels where they were associated with grit- and limestone-tempered pottery. The picture presented by the total distribution of the parallel-stemmed points was one of long usage for this general type. It would appear that from time to time distinctive modifications, as mentioned above, were made by certain groups. The most numerous points were those classified in the expanded-stem category. They made up 36.11% (65 points) of the total sample of 180 points. Of these 65 points, 43 (66.15%) were found in levels containing pottery and 22 (33.84%) were located in non-pottery levels. These points varied in outline but each possessed a stem with a slightly flaring base. The variation in size was considerable, but none was as large as some in the parallel-stemmed group. Some points of this general type could have been thrown into one of the notched-point categories but they would have represented poor examples of such types. It may be said that these points were characterized by their vagueness of form. The materials used in their manufacture showed great diversity. Expanded-stem points have been found at many sites in the Upper Ohio Valley. They are known from shell-midden sites along the Ohio River where they are associated with lanceolate and broad-stemmed types. They are most abundant on sites of Archaic times (Stewart and Dragoo, 1954) where they are the dominant type, but they are known to have persisted in use up through Middle Woodland occupations (Dragoo, 1954). Side-notched points. Points with side notches were second in abundance 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 161 % * 0 # # # 1 1 ? - ’■ • « ^ (i § * I s i i # #S#wt ‘ ‘ * ' , * , l : I#44«lf 1 1 • itiiliiiliu ** % !•« *# * k A i , # ■ * wj w Fig. 4. Dixon Shelter. Various types of points found during excavation 162 Annals of Carnegie Museum vol. 35 to those with expanded stems. The 58 points with side notches accounted for 32.21% of the sample of 180 points. During the sorting of this group it be- came apparent that there were two main categories in which the points could be placed. The first category consisted of points with varying out- lines and notches distributed throughout all levels in the excavation. The second category consisted of small, thin, slender points with deep, well formed notches and a limited level distribution. There were 39 points of a general side-notched type. See Fig. 4, no. 77-120 for examples. Of these points, 53.84% (21 points) were found in levels containing pottery while 46.15% (18 points) were found in levels below the pottery. Points of this type were found from the lowest to the top level but their greatest concentration occurred in levels three and four in blocks containing four to five levels and in levels two and three in blocks contain- ing only three levels. Many of these points are small and closely parallel the distribution of the expancled-stem type. Both of these types of point appear to belong to the same cultural tradition. It was often difficult to separate clearly the two types both in form and kinds of materials used in their manufacture. Both types are typical of sites now classified as of Archaic times in the Upper Ohio Valley. The second category of side-notched points contained 19 points with 78.94% (15 points) in pottery levels and 21.05% (4 points) in non-pottery levels. Their heaviest concentration occurred in levels two and three in blocks containing four or five levels and in level two in blocks having only three levels. The closest pottery association was with limestone-tempered ware. This type of point does not appear to be associated with any estab- lished Archaic manifestation in the Upper Ohio Valley. Its origin must be sought among later pottery-producing cultures. Corner-notched points. Points of this type do not compose a well de- fined category at the Dixon Shelter. Only seven points, or 3.88% of the total point sample, were tentatively assigned to the corner-notched type. Of these points 57.14% (four points) were found in levels containing pottery while 42.85% (three points) were below the pottery levels. These points stand apart from the expanded-stemmed type only by their deeper and more finely chipped notches, their greater width in relation to length, and their generally thinner cross-section. Triangular points. Small triangular points represented 7.77% (14 points) of the total sample. All of these points were found in levels containing pottery with 71.4% (10 points) recovered from the first level, 21.4% (three points) in the second level, and the remaining 7.2% (one point) in the third level. All of the points found in the first level were close to the surface and closely associated with the shell-tempered pottery. The small number of points found below the first level probably reached the lower depth in areas of disturbance. All of the triangular points are of types commonly associated with Late Prehistoric cultures. Drills Only 15 drills were found. They were distributed 6.7% (1 drill) in the first level, 33.3% (5 drills) in the second level, 26.7% (4 drills) in the third level, 20% in the fourth level, and 13.3% (2 drills) in the fifth level. Thus, 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 163 the second level contains the largest number of drills followed in diminish- ing order by the third, fourth, and fifth levels. In relationship to other materials, especially pottery, drills were more closely associated with lime- stone-tempered and grit-tempered pottery and with the points below the pottery levels than with shell-tempered pottery. None of the drills is large, the length ranging from 3.8 centimeters to 5.2 centimeters. Nine of the 15 drills have slightly expanded bases, five are straight, and the remaining one has a contracting stem. Those with slighly expanded bases were more common in the second and third levels while those with straight bases occurred more frequently in the fourth and fifth levels. The materials used in the manufacture of the drills varied, but fine grained cherts were most common. (See Fig. 5.) Scrapers Next to projectile points, scrapers were the second most abundant tool found in the excavation. They occurred in all levels and are of various types. Almost every flint chip found showed evidence of having been used as a scraper. For the purpose of classification only well worked pieces with obvious secondary chipping and evidence of repeated use were selected. The well defined scrapers were divided into three types— side-scrapers, end- scrapers, and gravers. Their characteristics and distribution were as follows: Side-scrapers. Scrapers possessing one or more straight cutting edges were common in all levels. They were usually made from irregular flakes of medium to fine grained cherts. A total of 60 specimens was assigned to this category. According to levels these 60 specimens were distributed 16.66% (10 scrapers) in the first level, 26.66% (16 scrapers) in the second level, 26.66% (16 scrapers) in the third level, 21.66% (13 scrapers) in the fourth level, and 8.33% (5 scrapers) in the fifth level. No basic distinction could be made for side scrapers occurring in pottery or non-pottery levels. End-scrapers. These scrapers are of the type commonly called “thumb- nail” scrapers. They are generally small and nicely chipped with a rounded cutting edge. Fine-grained cherts were most often used in their manufac- ture. End-scrapers were not too common and they had a more restricted distribution than the side-scrapers. The seven scrapers of this type were found 42.85% (3 scrapers) in the third level, 42.85% (3 scrapers) in the fourth level, and 14.28% (1 scraper) in the fifth level. End-scrapers were found in only the lowest pottery level and in the non-pottery levels. They would appear to be of definite Archaic context at this site. (See Fig. 5.) Gravers. Only three objects were assigned to this category. They were small with triangular outline and cross-section. They all had finely chipped, sharp points and one or two sides worked as scraper edges. They were made of fine-grained chert. Two of these tools were found in the second level and one in the fourth level. BONE ARTIFACTS Animal bones of a refuse nature were found scattered throughout the various levels of the site. John Guilday of the Museum staff has identified these bones in the following report. 164 Annals of Carnegie Museum vol. 35 Fig. 5. Dixon Shelter. Points, drills, scrapers, bone tools, and copper beads found during excavation 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 165 Report on the bone refuse from the Dixon Rock Shelter (46 Pr 6) The bone refuse, 2467 fragments in all, consisted of: unidentified mammal fragments, 2116; identified mammal fragments, 332; bird 8; turtle 4; and one fragment of human bone. Ursus americanus Black bear: 4 bone fragments. Procyon lotor Raccoon: 3 bone fragments. Marmota monax Woodchuck: 2 bone fragments. Tamias striatus Chipmunk: 1 bone fragment. Sciurus carolinensis Gray squirrel: 3 bone fragments. Castor canadensis Beaver: 1 tooth. - Erethizon dorsatum Porcupine: 3 bone fragments. Sylvilagus sp. Rabbit: 1 bone fragment. Ceruus canadensis Elk: 30 bone fragments. Odocoileus virginianus White-tailed deer: 283 bone fragments. Unidentified mammal bone: 2116 fragments. Colinus virginianus Quail: 1 bone fragment. Meleagris gallopavo Wild turkey: 4 bone fragments. Bird sp.: 3 bone fragments. Turtle sp.: 4 bone fragments. Man: One phalange, infant. One fragment of turtle carapace showed signs of “sanding” on both external and internal surfaces. Possibly a cup fragment. Only a few tools made from bone were found in the excavation. None could be called distinctive for any particular time period. Most common of the bone tools were antler tip points, antler drifts or flakers, splinter bone awls, and beads. Nearly all of these objects were broken, and some had been partially burned. Two of the antler tips had been intentionally fire hard- ened. All of the awls were made from crude splinters of bone with only the tips carefully worked. The antler drifts showed more careful attention. In Fig. 5, no. 111-122, the best of these items are illustrated. One deer phalange had been modified into an artifact. The end had been pierced and grooved. It was probably used as a tinkler. A fragment of turtle carapace showed signs of sanding on both external and internal surfaces. It was possibly a cup fragment. The only unusual item of bone was a distal joint of a deer canon bone 1% inches in length which had been drilled for suspension. (See Fig. 5, no. 110.) The preservation of bone at the Dixon Shelter was not good. Bones found in the lower levels were very fragmentary and bones in the upper level were poorly preserved. The sample of worked objects does little to clarify the cultural picture of the Dixon Shelter. 166 Annals of Carnegie Museum vol. 35 METAL ARTIFACTS With the exception of a historic pocket-knife and an iron ring found near the surface only two objects of metal were found in the excavation. Both of these objects are rolled copper beads from the third level in division A. These beads were found a few inches apart and in association with grit- tempered and limestone-tempered pottery. DISCUSSION The Dixon Shelter was occasionally occupied from Archaic times until Late Prehistoric times. All major cultural time periods with the exception of Paleo-Indian were represented by materials found in the shelter. Approxi- mately one-third of the debris deposit of the shelter could be assigned to the non-ceramic Archaic period while the upper two-thirds of the deposit con- tained ceramic materials of the three major ceramic culture periods— Early Woodland, Middle Woodland and Late Prehistoric. The Dixon Shelter was a convenient station for small hunting parties. It afforded shelter from rain and strong winds and was adjacent to fresh water. Completely surrounding the shelter was an area containing a good population of game. In Archaic times, there were no large groups of people in the Upper Ohio Valley. There were probably numerous small family groups inhabit- ing, where available, small rock shelters. In later ceramic times, after the introduction of agriculture, larger populations established permanent vil- lages and rock shelters became stations for hunting parties and families on collecting trips. The materials from the Dixon Shelter present an interesting study of the transition from Archaic into the later ceramic periods. The tool assemblage of Archaic levels was characterized by the use of small notched and stemmed points, side- and end-scrapers, and crude knives. The use of small notched points began at the lowest level and continued in use up into the Early Woodland and Middle Woodland periods. Large stemmed points were limited to the upper levels of the Archaic deposit and were found occasion- ally in the later culture periods. With the advent of heavy, grit-tempered pottery came the introduction of the distinctive parallel-stemmed point with rounded base. This type of point was commonly associated with the Adena culture in the central Ohio Valley and it appears to be a good time marker in the Upper Ohio Valley. At this time level there appears a preference for finer materials for the manu- facture of points. Some of these materials were imported from areas out- side the Upper Ohio Valley, particularly from the Flint Ridge deposits in Ohio. The major change that appears during the Middle Woodland period is the use of limestone-tempered pottery. There are no point types different from those prior to that time. The medium-thick grit-tempered pottery also remained in use. The most abrupt change occurs with the advent of the Late Prehistoric period. Major changes occurred in types of pottery and projectile points. 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 167 The pottery became predominantly shell tempered and projectile points were nearly always of the small triangular type. Influences from the Potomac River area were at a minimum at the Dixon Shelter. Only a few specimens were of materials generally associated with coastal complexes, and some of this material is known from the local area. The materials used in the manufacture of tools at the Dixon Shelter con- form to those commonly found at sites in the Upper Ohio Valley. Most of these materials are of local origin occurring either in pebbles or thin vein deposits. A wide variety of materials was used, some of which were poorly adapted to the Indian’s needs. The most distinctive import was small quan- tities of Flint Ridge chert. From the standpoint of materials, cultural movements spread up the Ohio Valley and into the East. The distribution of points in the Dixon Shelter is indicated in Tables 1-3. TABLE 1. DIXON SHELTER, DIVISION A. DISTRIBUTION OF POINTS IN BLOCKS HAVING FIVE LEVELS Level T CN ST EXP-ST SNT SN M Total 1 5 3 2 1 11 2 2 1 1 3 4 5 1 17 3 4 4 5 1 14 4 3 1 2 3 9 5 2 1 3 Total 7 1 13 8 9 9 7 53 T = T riangular; CN — Corner-notched; ST = Stemmed; EXP-ST = Expanded-stem; SNT = Side-notched thin; SN = Side-notched; M = Miscellaneous. TABLE 2. DIXON SHELTER, DIVISION A. DISTRIBUTION OF POINTS IN BLOCKS ] HAVING THREE LEVELS Level T CN ST EXP-ST SNT SN M Total 1 4 7 1 12 2 7 8 3 4 1 23 3 2 3 7 6 18 Total 0 2 14 22 3 11 1 53 T = Triangular; CN = Corner-notched; ST = Stemmed; EXP-ST : = Expanded-stem; SNT = Side-notched thin; SN = Side-notched; M = Miscellaneous. TABLE 3. DIXON SHELTER, DIVISIONS B AND C. DISTRIBUTION OF POINTS BY LEVELS Level T CN ST EXP-ST SNT SN Total 1 5 2 5 2 1 1 16 2 1 1 4 8 1 1 16 3 1 11 1 7 20 4 1 10 4 9 24 5 4 1 5 Total 7 4 9 35 7 19 81 T = Triangular; . CN = Corner- notched; ST — Stemmed; EXP-ST — Expanded-stem; SNT = Side-notched thin; SN = Side-notched. 168 Annals of Carnegie Museum vol. 35 THE ROHR ROCK SHELTER During our excavation at the Dixon Shelter in August 1954, Albert Bauer took me to a fine rock shelter close to the Cheat River near Rohr in Monon- galia County, W. Va. This site was named the Rohr Shelter and given the number 46 Mg 9 in the files of Carnegie Museum’s Upper Ohio Valley Archeological Survey. By examining the surface and digging a small test pit, we determined that this site contained cultural materials similar to those at Dixon. The test also indicated greater depth to the deposit and a heavier concentration of Archaic materials. This site presented an excellent opportunity to cross check our findings at Dixon and to extend our knowl- edge of the Archaic in time. With the kind co-operation and permission of Edward Howard, the owner of the site, plans were made for the excavation of the shelter in 1955. On May 23, 1955, work was begun with the aid of Mr. and Mrs. Albert Bauer. On May 28, John Leppla and Mrs. Dorothy Dragoo joined the crew. Major excavation work was completed on June 6, but smaller areas have been excavated since that time by Bauer and Leppla who have given much of their time and energy to the project and without whose aid the work would not have been completed. NATURAL SETTING The Rohr Shelter is located in low mountainous country at an elevation of 2025 feet above sea-level. The surrounding hilltops reach an elevation between 2100 and 2300 feet. The shelter is in a narrow valley through which a small stream flows into the Cheat River about one mile to the east. This stream now flows nearly fifty feet below the floor level of the shelter. The overhanging ledge or roof of the shelter is composed of Pottsville conglomerate, the basal formation of the Pennsylvanian period in the area. It is a compact, grayish white sandstone with a dense inclusion of smooth quartz pebbles. Underlying the conglomerate is a fine, brown sandstone which has been eroded away to form the shelter. All of the major shelters I have seen in the Cheat River area have been composed of these same formations. The vegetation is very much like that at the Dixon Shelter, and the animal life in Prehistoric times was probably quite similar. EXCAVATION The overhanging roof of the Rohr Shelter extends 92 feet along the out- crop in a line from northwest to southeast. From the front of the shelter to the rear wall, the recess varies in depth from a few feet to as much as 45 feet. At the front of the shelter the ceiling ranges in height from three to six feet above the present floor level. The ceiling slopes gently downward toward the rear of the shelter in an area varying in depth from 10 to 15 feet; and then lowers abruptly to a height of three feet which then lowers gently to the rear of the shelter. Near the northwestern margin of the shelter is a spring and an area of intermittent water flow from under the outcrop. No cultural deposits were found in this area. Only a few stones had fallen from the ceiling. All appear to be old falls as they were found to be resting on 1958 Bragoo: Archaic Hunters of the Upper Ohio Valley 169 the sterile floor and surrounded by the cultural bearing layers. On the ground surface directly below the outer margin of the shelter was a ridge of materials deposited by water action from the hill slope above the shelter. Directly in front of the shelter was a dense growth of trees and mountain laurel. The area of the shelter was surveyed and a base-line was established parallel with the front of the ceiling. From this base-line, a grid of blocks 10 feet square was marked for the entire habitable area of the shelter. A datum point was established for the taking of elevations and the drawing of profiles. Each block in the grid was excavated in six-inch levels beginning at the surface and numbered in sequence from top to bottom. All the materials from each level of each block were bagged separately and catalogued. Any materials from a feature within a block were catalogued separately from materials at the general level. The excavating was done by customary archeological methods. Trowels were used for nearly all primary digging. Most objects were found during the removal of the soil by trowel, but because of the general darkness of the shelter it was necessary to screen all the soil to prevent loss of small items. Observations of soil differences, features, and cultural materials were made for each level within a block. Drawings and photographs were made of all important features. Initial excavation was concentrated in five blocks 10 feet square, running in a line parallel with the front of the shelter and including the most habitable area. The cultural debris in this area, 50 feet in length, varied in depth from 4*4 feet near the center of the shelter in blocks two, three, and four to a minimum depth of two feet in block one and three feet in block five. (See Fig. 6.) The debris layer was composed of three distinct layers of soil. The first and topmost layer extended from the surface to a depth of approximately six inches. The soil of this layer was gray. Most of the Late Prehistoric ma- terials were found within the confines of this layer. Immediately below the layer of gray soil was a layer of dark soil with scattered inclusions of yellow sand. This layer varied in depth from 6 inches near the ends of the shelter to 18 Inches at the center of the shelter. Throughout this layer pottery sherds of the Early Woodland and Middle Woodland periods were found. Below the layer of dark soil was a layer of very black soil. Scattered through- out this layer were remains of many fires which had in many places turned the soil a brick red. This layer varied in depth from one foot near the ends of the shelter to 2]/> feet at the center. This very black soil was devoid of pottery sherds. On the basis of the stone materials recovered, this layer represents the major Archaic occupation of the shelter. (See Fig. 7.) Below the very black soil was a yellow sand which appeared free from archeological materials except for a few objects which occurred on or just below its surface. Several tests dug into this yellow sand produced no evi- dence of cultural materials or disturbance by man. However, in one small test pit in block two a point and several flint chips were found in an area 170 Annals of Carnegie Museum vol. 35 of reddish sand nearly six inches below the surface of the yellow sand. The full extent of this reddish stained sand is yet to be determined. The evi- dence is inconclusive as to whether the red sand, with the artifacts found in it, was intrusive into the yellow sand or represents a very old occupation of limited extent. It is hoped that future work will solve this problem. Fig. 6. Rohr Shelter. General view of excavations with Albert Bauer re- cording the features There were no major features in the uppermost layer composed of gray soil. A few thin areas containing darkened soil and small pieces of char- coal were noted. No definite fire-pits were found as the fires had not been confined to a depression in the soil but had been placed directly on the surface. Around these areas of fire were a thin accumulation of ash, small bits of charcoal, and slightly reddened earth. Only fires of short duration were represented by these areas in the gray-soil layer. This situation would lend support to the idea that rock shelters received only transitory use during Late Prehistoric times. A few accumulations of debris were found in the gray soil but no well defined pits were discovered. The evidence indicates that refuse was de- posited on the surface without benefit of pit disposal. Beginning at approximately six inches below the surface the layer of dark soil presented features of a well defined nature. Nine stone lined fire- pits were found within the layer of dark soil from 6 to 18 inches below the surface. All of these fire areas were similar in construction. Three to four flat stones had been placed in a slight depression in the soil and then a 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 171 ring of stones placed around them. These fire-pits presented evidence of long and intensive use as there were considerable concentrations of charcoal and ash. The stones were cracked and discolored and the soil around the stones was burned a brick red. None of the f. re-pits contained any quantities of cultural materials. An occasional flint chip or point was found within them. Several small areas, discolored by fire but not outlined by stones, were also found in the dark layer. As with similar features found in the gray soil above, these areas were thin and poorly defined. Well defined refuse pits were also absent from the dark soil layer. There were, however, concentrations of refuse thinly spread over small areas. The very black soil immediately below the dark soil and above the yellow sand contained a number of stone-lined fire-pits similar to those found in the layer of dark soil. Fires had also been kindled among the few large boulders which protruded several inches above the surface of the yellow sand. The entire black soil layer, varying in thickness from 1 to 214 feet, contained very extensive evidence of many fires. The black coloration of the soil was caused by large accumulaions of charcoal and ash. Almost every small stone found within the layer was discolored and cracked by heat. The soil in several areas was burned a brick red. Fig. 7. Rohr Shelter. Hearth at 18-inch level The layer of black soil presented a picture of long occupation. Since the cultural materials found in the black soil were of Archaic times, it is ap- parent that the shelter provided an ideal home for small hunting and col- 172 Annals of Carnegie Museum vol. 35 lecting bands of people. The shelter would have been at its maximum utility during Archaic times as the ceiling would have been from two to nearly four feet higher from the floor than it is at present; as the refuse and soil accumulated, the height of the shelter gradually decreased. As well as the diminishing desirability, in later times the changed economy and village life of peoples living in the region made the shelter of only occasional use. (See Fig. 8.) POTTERY Pottery was not abundant in the refuse of the Rohr Shelter but a suf- ficient number of sherds was found to give a general picture of the sequence and the types involved. All information is based upon the study of sherds since no whole or completely restorable vessels were found. All sherds from the levels of each block were sorted and classified ac- cording to “wares” based on the kind of temper. Four major “wares”, as defined in Carnegie Museum’s Upper Ohio Valley Archeological Survey’s nomenclature, were represented— shell-tempered Monongahela, limestone- tempered Watson, medium grit-tempered Mahoning, and coarse grit-tempered Half-Moon. These are the same wares found at the Dixon Shelter and their full cultural relationships were discussed in the previous section. Each of the “ware” groups found at Rohr Shelter was sorted into its component types with the following results. Monongahela Cordmarked This type is manufactured by coiling and the use of a cordwrapped paddle. The temper is moderately to finely crushed particles of shell com- posing approximately 25% of the paste. The texture is finely laminated and the hardness ranges from 2.5 to 3.5. The color on the exterior surface ranges from buff to black while interiors usually are gray to black. The exterior is covered by the impressions of a cordwrapped paddle. These impressions usually meet the lip at an oblique angle. On some sherds the cord impres- sions have been smoothed. The interior surface is smooth. Decoration is rare; when present, it is generally confined to the lip, and consists of small indentations of a cordwrapped stick or the edge of a paddle. The most common forms of this type as known from other sites are globular jars and bowls with flattened or rounded lips on a straight or slightly flaring rim. Thickness ranges from 3.5 to 7 mm. Only 65 sherds of this type were found during the excavation of the Rohr Shelter. Monongahela Plain This pottery is identical with Monongahela Cordmarked with the excep- tion of having a smooth, plain surface. Only 10 sherds of this type were found in the Rohr sample. Watson Cordmarked Pottery of this type was manufactured by coiling and malleating with a cordwrapped paddle. Crushed fragments of limestone making up from 25 to 50% of the paste were used for temper. The texture is compact but often irregular because of the large size and angular shape of the limestone temper. The hardness varies from 2.5 to 3.5. The color of both interior and exterior 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 173 surfaces is usually a dull gray or tan. The exterior surface is impressed with the marks of a cordwrapped paddle. The interiors are generally smooth but occasionally faint cord impressions occur. Decoration is rare and confined to the lip or neck. The common form for this vessel is globular with slightly flaring rim. Thickness ranges from four to eight mm. The Rohr sample in- cludes a total of 130 sherds of this type. Watson Plain This type is similar in all features to Watson Cordmarked except for the smooth surface finish. Eight sherds are in the sample from Rohr. Mahoning Cordmarked Pottery of this type was manufactured by coiling and the use of the cordwrapped paddle. The temper is crushed fragments of igneous rock, chert, or quartz of moderate to small size which make up 25 to 40% of the paste. The texture is compact and regular and the hardness ranges from 2.5 to 3.5. The color ranges from dark gray to tan. Surface finish consists of 174 Annals of Carnegie Museum vol. 35 cord impressions on the exterior while the interior surface is usually plain. Decoration is confined to the lip and consists of the impressions of a cord- wrapped paddle. The form was probably globular with a slightly flaring rim. The thickness ranges from 4 to 7 mm. Only seven sherds of this type are in the Rohr sample. Mahoning Plain Mahoning Plain shares all the general features of Mahoning Cordmarked except for the smooth surface. The geographic and cultural relationships of the two types appear identical. There were 1 1 sherds of Mahoning Plain in the Rohr sample. Half-Moon Cordmarked Pottery of this type was manufactured by coiling and malleating with a cordwrapped paddle. The temper is coarsely crushed fragments of igneous rock, chert or quartz. A combination of these tempering materials also occurs. Temper fragments are large and make up from 50 to 70% of the paste. The texture is rough and irregular and the hardness ranges from 2.5 to 3.5. The color varies from tan to light gray. The exterior surface is deeply marked by cord impressions. The interior surface is usually marked by cord impressions but plain interiors also occur. No decoration occurred on this type at Rohr. The form was apparently large globular vessels with some vessels having flat bases. The thickness ranges from 8 to 16 mm. POTTERY DISTRIBUTION A total of 239 pottery sherds of all types was removed from the levels of blocks A1 to A5 which constituted an area of 500 square feet and constituted the main habitable area of the shelter. Of the 239 sherds from this area, 192 or 80.3% were found in the first six-inch level, 37 sherds or 15.4% in the second level, 9 sherds or 3.8% in the third level, and 1 sherd or 0.4% in the fourth level. Thus, 95.8% of all pottery found was concentrated within the first two levels. The few sherds found in level three and the single sherd from the fourth level may well represent sherds displaced from the upper levels through disturbances. The first-level pottery sample of 192 sherds contained examples of all the major “wares” known from the site. Monongahela ware made up 33.33% (64 sherds) of the sample. Watson ware was most prominent with 57.29% (109 sherds). Mahoning ware composed 6.77% (13 sherds) of the sample. Sparsely represented was 1.04% (2 sherds) of Half-Moon ware. The remain- ing 2.08% (4 sherds) were unidentifiable because of their smallness and fragile condition. Of the 37 sherds recovered from the second level 29.72% (1 1 sherds) were Monongahela ware, 62.16% (23 sherds) Watson ware, and 8.11% (3 sherds) Mahoning ware. No Half-Moon ware was found in this level. Only nine sherds were found in the third level. Watson ware accounted ior 55.5% (5 sherds) of the sample. The remaining sherds were divided evenly 22.2% (2 sherds) Mahoning ware and 22.2% (2 sherds) Half-Moon ware. No shell-tempered pottery was found in this level. One sherd of Watson ware was found in the fourth level of block A2. This was the only sherd found below 18 inches in the entire excavation. 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 175 All sherds in A1 and A5 were within the first 12 inches and those in A3 and A4 were contained within the first 18 inches of the deposit. The shell-tempered pottery ware consisted of two major types— Monon- gahela Cordmarked and Monongahela Plain. Of the total 75 shell-tempered sherds 86.6% (65 sherds) were cordmarked and the remaining 13.3% (10 sherds) had a plain surface. The Monongahela Cordmarked type was dis- tributed 84.6% (55 sherds) in the first level and 15.3% (10 sherds) in the second level. Of the Monongahela Plain sherds 90% (9 sherds) were in the first level and 10% (1 sherd) in the second level. Of the entire sample of shell-tempered pottery, 85.3% (67 sherds) were found in the first level and 14.6% (11 sherds) in the second. From observations made during the excava- tion and the sherd tabulations it was apparent that shell-tempered pottery was within the last phase of pottery making at the shelter. Many of the sherds recorded for the first level were found on or near the surface, and those found in the second level were from blocks containing the greatest depth of debris. The limestone-tempered ware was of two types, Watson Cordmarked and Watson Plain. Of the 138 limestone-tempered sherds 94.2% (130 sherds) were cordmarked and 5.8% (8 sherds) were plain. Limestone-tempered pot- tery was found from the first to the fourth level and constituted the most abundant and longest pottery sequence found in the excavation. The 138 sherds of this sample were distributed 78.9% (109 sherds) in the first level, 16.6% (23 sherds) in the second level, 3.6% (5 sherds) in the third level, and 0.7% (1 sherd) in the fourth level. Watson Cordmarked and Watson Plain occurred together in all levels except the fourth which contained the single cordmarked sherd. As with the shell-tempered pottery, the limestone- tempered pottery shows its greatest concentration in the first level, but it begins two levels below the first appearance of shell-tempered pottery and is representative of Hopewellian and other so-called Middle Woodland mani- festations. In deeply stratified sites it is always below the shell-tempered types so common to Late Prehistoric times. Because of the greatest abundance of pottery in the first level at the Rohr Shelter, it is difficult to separate clearly limestone-tempered and shell-tempered pottery in time. The super- ficial location of many shell-tempered sherds with triangular points would suggest that the two wares were not contemporaneous and may not have influenced each other in development. The medium grit-tempered pottery was of two types— Mahoning Cord- marked and Mahoning Plain. Of the sample of 18 sherds 38.8% (7 sherds) were cordmarked and 61.1% (11 sherds) were plain. All of the cordmarked sherds were found in the first level, but the plain sherds were distributed 54.5% (6 sherds) in the first level, 27.2% (3 sherds) in the second level, and 18.2% (2 sherds) in the third level. The distribution of the total medium grit pottery was 72.2% (13 sherds) in the first level, 16.6% (3 sherds) in the second level, and 11.1% (2 sherds) in the third level. The medium grit- tempered pottery appears to have been an outgrowth of the early heavy grit-tempered pottery tradition in the Upper Ohio Valley. It is found as a minority ware on most sites which produce limestone-tempered pottery which also appears to be related to the early heavy grit-tempered tradition. 176 Annals of Carnegie Museum vol. 35 The heavy grit-tempered ware known as Half-Moon Cordmarked in the Upper Ohio Valley is represented by only four sherds taken from the excava- tion in the Rohr Shelter. Two of these sherds were found in the first level and the other two in the third. Since sherds of this type are found only at the lowest level in deeply stratified sites, the finding of two such sherds in the first level at the Rohr Shelter could possibly indicate a disturbance not noted during the excavation. The general picture presented by the distribution of pottery at Rohr was one of compactness. Although types representative of all major ceramic periods were present, the greatest number of sherds for each type was found in the first level. Below the first level the number of sherds dimin- ished rapidly. The time span represented by all the pottery types found within the first three levels would account for nearly 2500 years during three major cultural periods. Below the pottery-bearing levels were three to five levels, depending upon area, of cultural debris containing artifacts from much earlier times. CHIPPED-STONE ARTIFACTS Projectile points The most abundant chipped-stone artifacts recovered from the excava- tion were projectile points, with a total of 288 points collected from all blocks. Projectile points were found in all six of the debris-producing levels of the excavation. Their distribution according to levels was 78 (27.08%) in the first, 54 (18.75%) in the second, 75 (26.04%) in the third, 60 (20.83%) in the fourth, 14 (4.86%) in the fifth, and 7 (2.43%) in the sixth. All of the points were sorted into generally accepted types and plotted as to distribution by levels. It was difficult to differentiate clearly many of the small expanded-stem and side-notched points because of their crude- ness and varied form. Many of these points possessed characteristics of both types. As will be seen later in the discussion of these points and their dis- tribution, the attempt to divide them into two types may be more artificial than real. Distribution is indicated on page 186. In the discussion which follows, each type of point and its variations will be described, and its relationship to the pottery distribution noted. (See Fig. 9-14.) Triangular points. The 27 small triangular points represented 9.37% of the total sample. All of these points were confined to the first two levels, with the first level containing 26 points (96.29%) and the second level having only 1 point (3.70%). In the first level there were 78 points of various types. Triangular points were the dominant type, constituting 33.33% of the 78 points. Almost all of the triangular points were found near the surface and closely associated with the shell-tempered pottery. The one point found in the second level probably reached the lower depth through disturbance of the soil. All of the triangular points are of the type commonly associated with Late Prehistoric culture in the Upper Ohio Valley. Corner-notched points. The points of this type do not compose a well defined category at the Rohr Shelter. Only six points or 2.08% of the total point sample were assigned to this group. Their distribution was one 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 177 Fig. 9. Rohr Shelter. Artifacts from level 1 (0"-6") Fig. 10. Rohr Shelter. Artifacts from level 2 (6"- 12") 178 Annals of Carnegie Museum vol. 35 Fig. 11. Rohr Shelter. Artifacts from level 3 (12"-18") Fig. 12. Rohr Shelter. Artifacts from level 4 (18"-24") •II 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 179 F.g. 14. Rohr Shelter. Artifacts from level 6 (30"-36") in row 1-3, and from level 7 (36"-42") in row 4 180 Annals of Carnegie Museum vol. 35 point (16.66%) in the first level, three points (50.00%) in the second level, and two points (33.33%) in the third level. All of these points came from levels which produced pottery. Corner-notched points stand apart from the expanded-stem type only by their deeper and more finely chipped notches, their greater width in relation to length, and their generally thinner cross-section. Points such as these are similar to those known for late Early Woodland and early Middle Woodland manifestations. Stemmed, points. All the points assigned to this type possessed well de- fined stems with parallel sides. The bases of the stems were either flat or rounded. The majority of these points were four centimeters or more in length and were the largest points found in the excavation. Most of the stemmed points were made from crude materials which could not be finely chipped. From looking at these points one could easily get the impression that they should be old, but their distribution does not support this view. Of the total 19 points of this type, nine points (47.39%) were found in the first level, nine points (47.39%) in the second level, none in the third level, and one (5.26%) in the fourth level. Only the point in the fourth level was found in an area not containing pottery. Three of these stemmed points with rounded bases and made of fine-grained chert are similar to those associated with the Adena culture. The remaining 16 points appear closely related typologically to some Ohio Valley late Archaic complexes. Large expanded-stem points. During the sorting of the expanded-stem points it became apparent that the large points of this group were found more commonly in the upper levels of the excavation in contrast to the greater depth distribution of small points with expanded-stems. Any point of the expanded-stem group four centimeters or more in length was put into the large expanded-stem type and all those below four centimeters in length were classified as the small expanded-stem type. The large points were more crude and usually made of a poorer variety of flint than the small points. The 10 points of the large expanded-stem type were distributed six points (60%) in the first level, one point (10%) in the second level, two points (20%) in the third level, and one point (10%) in the fourth level. Seven of these points were found in levels containing no pottery. The one point found in the fourth level was unusual in that it had broad serrations along the cutting edges. The large stemmed points closely approximate the distribution of the parallel-stemmed points and are probably culturally related. They are also part of a long tradition of expanded point making which began early in Archaic times in the Upper Ohio Valley. Small expanded-stem points , Of the 288 points found in the excavation, 134 or 46.52% were of the small expanded-stem type. All of the points in this category were less than four centimeters in length. A great variety of flints was used in their manufacture, but generally the material was finer textured than that used for the large expanded-stem points. Small stemmed points were distributed throughout all levels from the first to the sixth with 22 (16.41%) in the first, 26 (19.4%) in the second, 46 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 181 (34.32%) in the third, 29 (21.64%) in the fourth, 9 (6.71%) in the fifth, and 2 (1.49%) in the sixth. They are the dominant type of point in all levels except the first, where triangular points are more abundant. In the Upper Ohio Valley small stemmed points are found as the major type of point on the majority of Archaic sites where they appear to have remained in use over a long period of time. Large side-notched, points. The side-notched sample of points was divided into two groups according to size. Any point four centimeters or longer was considered large while any point less than four centimeters was small. Un- like the distribution of the large expanded-stemmed point, which was con- fined to the upper levels, the large side-notched point was found through- out all levels. Of the 20 points of this type there were five (25%) in the first level, three (15%) in the second level, two (10%) in the third level, eight (40%) in the fourth level, one (5%) in the fifth level and one (5%) in the sixth level. As was true of the large expanded-stem points, many of the large side- notched points were of coarse material and not nearly so finely chipped as the small type. The one large-stemmed point found in the sixth level was unique because of the fine flint of which it was made and the excellent workmanship. Small side-notched points. Small side-notched points were second in abun- dance at the Rohr Shelter. The sample of 65 points represented 22.56% of the total 288 points. Small side-notched points were found in all point- producing levels and presented the following distribution. Eight points (12.3%) in the first level, 10 points (15.38%) in the second level, 22 points (33.84%) in the third level, 19 points (29.23%) in the fourth level, 4 points (6.15%) in the fifth level, and 2 points (3.07%) in the sixth level. The small side-notched points and the small expanded-stem points had a number of common features. Both types were wide in ratio to length and usually rather crudely chipped. During the sorting of the points it was often difficult to decide clearly in which type many of these points should be placed. A wide variety of materials was used in the making of both types. Bifurcated points. Only three points assigned to this type were found at the Rohr Shelter. Such points are small, less than three centimeters long, and relatively broad with a deeply notched base and well defined shoulders. All of these points were found in the lower levels and were not associated with pottery. Their distribution was as follows. One point (33.33%) in the fourth level, and two points (66.66%) in the sixth level. In the Upper Ohio Valley, bifurcated points occur in Archaic manifesta- tions. As yet, their distribution is not well known as they occur only sparsely on all sites from which they have been reported. Miscellaneous points. Four points found during the excavations could not be placed in the types just described. In the first level was found a tapered-stem point which was quite differ- ent from any belonging to the stemmed-point type. From weakly formed shoulders the stem tapers abruptly to the base which is straight. The material of which this point was made is a crude chert which could not be finely chipped. 182 Annals of Carnegie Museum vol. 35 The second point of interest in the miscellaneous group was a small lozenge-shaped point found in the first level. The basal portion of this point makes up approximately one-half the total length of the point. It was 3 centimeters in length and 2.8 centimeters in width. The base is slightly rounded. In the third level was found a finely chipped lanceolate blade 4.5 centi- meters in length and 1.9 centimeters in width. This point has a thinner cross-section than any of the stemmed or notched points found in the ex- cavation. The base is slightly rounded. The material of which the point was made is a reddish, soft chert which has been weathered. A crude lanceolate point 3.8 centimeters in length and 2 centimeters in width was found in the fourth level. The sides of this point are parallel and the base is straight. The material of this point is a coarse-grained chert. There was no pottery associated with this point. Projectile-point summary. The various types of projectile points and their distribution at the Rohr Shelter present some interesting insights into the cultural sequences of the Upper Ohio Valley. When information per- taining to the point types is combined with that of the pottery types and their distribution, there emerges evidence for four major cultural complexes— Late Prehistoric, Middle Woodland, Early Woodland, and Archaic. The small triangular points were found only in the uppermost levels and usually on or very near the surface. Such points are well known from all Late Prehistoric sites throughout the Ohio Valley. Their association with shell-tempered pottery adequately demonstrates their cultural context at the Rohr Shelter. Corner-notched points were represented by only a few specimens. All of these points were confined to levels containing pottery of predominantly limestone or grit temper. Culturally they are generally considered to be of late Early Woodland and Middle Woodland origin. Points with parallel stems were basically concentrated within the first and second levels where they were closely associated with grit-tempered and lime- stone-tempered pottery. A few of the stemmed points with rounded bases and made of fine-grained flint are similar to those known from Adena sites. At the Rohr Shelter stemmed points probably were first used during late Archaic times. Large expanded-stem points were also concentrated within the upper- most levels where pottery occurred, but a few of these points were found in the upper non-pottery levels where they appear to be part of a late Archaic complex. In contrast to the large expanded-stem points, and parallel-stemmed points, large side-notched points were found in all levels and were equally divided between pottery and non-pottery levels. They appear to be the large counterpart of the small side-notched points so common on many Archaic sites and found in all levels at the Rohr Shelter. Small expanded-stem and small side-notched points constituted 69.08% of the entire point sample from the excavation. Both of these types were found in all levels. The small expanded-stem point was the dominant type in 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 183 all levels except the first where the triangular type was more abundant. The small side-notched point was second in abundance in all levels except the first. Small points were basic items in the tool complex of many Archaic sites in the Upper Ohio Valley. At the Rohr Shelter there is evidence that they were used for a long time before the introduction of pottery and that they persisted for some time after new cultures had entered the area. The small bifurcated points were too few in number to be a reliable sample, but their placement in the lowest levels would indicate that they also belonged to an early Archaic tradition. Scrapers and drills There were a number of scrapers, drills, and other chipped-stone objects found throughout the various levels of the excavation. An attempt was made to classify these objects in generally accepted categories. In the follow- ing paragraphs each of these categories will be described and the distribution of such objects according to levels will be noted. Side-scrapers. There were 68 side-scrapers found throughout the first six levels of the excavation. The distribution was 6 (8.82%) in the first level, 8 (11.76%) in the second level, 18 (26.47%) in the third level, 23 (33.82%) in the fourth level, 2 (2.94%) in the fifth level, and 11 (16.17%) in the sixth level. Nearly 79.42% of the side-scrapers were found in levels which did not contain pottery. As has just been noted, the greatest con- centrations of side-scrapers occurred in the third, fourth, and sixth levels which are considered to be Archaic in time. Side-scrapers varied in shape but all possessed one or two straight cutting edges which in the majority of the examples showed secondary chipping or retouching. Side-scrapers were used by all peoples who inhabited the Rohr Shelter from early to late times, but they appear to have been a more important item in the tool kit during Archaic times. End-scrapers. There were 25 end-scrapers found in the excavation. All of these scrapers were found below the upper two levels in which pottery was found in abundance. Their distribution was 5 (20%) in the third level, 6 (24%) in the fourth level, 12 (48%) in the fifth level, and 2 (8%) in the sixth level. This distribution would indicate a definite Archaic context for end-scrapers at the Rohr Shelter. All of these scrapers were well made and exhibit in most examples fine secondary chipping. None of the end- scrapers was large and many of them were very small. They ranged in length from 3.5 centimeters to 1.5 centimeters. The material used for their construction was usually a fine-grained chert. This contrasted with the wide range of materials used for side-scrapers. Crude scrapers.. This category included many small pieces of flint which probably had been used only once and then discarded. Such pieces lacked the definite form and secondary chipping common to the side- and end-scraper categories. Nearly every piece of flint removed from the excavation showed in- dications of use. There were 480 such pieces with the distribution 63 (15.44%) in the first level, 83 (17.29%) in the second level, 96 (20%) in the third level, 105 (21.87%) in the fourth level, 105 (21.87%) in the fifth level, 21 (4.37%) in the sixth level, and 7 (1.45%) in the seventh level. Such items were thus distributed throughout all levels with the greatest number occurring in the 184 Annals of Carnegie Museum vol. 35 fourth and fifth levels. As time and cultural markers these crude scrapers were of little value. Miscellaneous scrapers. There were seven specialized scrapers of various types which could not be included in the above categories. In the first level was found one small scraper shaped in general outline like an end-scraper but with a concave, gouge-like cutting edge. One stemmed, bunt-like scraper was found in the second level and one similar object was recovered from the fourth level. In the fifth level were found two small gravers made of fine-grained flint with triangular cross-section and sharp, cutting points. In the seventh and lowest level producing artifacts an interesting chipped stone gouge was found. This tool measured 3.2 centimeters in length by 1.8 centimeters in width. It was finely chipped from a light gray chert. The lower surface was deeply grooved to produce a concave cutting edge. Also from the seventh level was a crude chopper-like tool with evidence of having been used as a scraper. Drills. Thirteen drills or fragments of drills were found. All of these drills appear to be either of the straight or slightly expanded base varieties. Their distribution was five (38.46%) in the first level, four (30.76%) in the second level, one (7.69%) in the third level, one (7.69%) in the fourth level, and two (15.38%) in the fifth level. No drills were found in the sixth and seventh levels, and 69.22% of them were found within the upper two levels containing pottery. Unidentified chipped-stone artifacts. There were 26 pieces of chipped stone which could not be assigned to specific categories. Most of the objects in this group were point tips or other portions of points not large enough to permit definite placement as to type. Such fragments were found through- out the first to the sixth level. Polished-stone artifacts Only a few items of polished stone were found in the excavation. A few hammerstones were found near the fire-pits in the third and fourth levels. Two pitted stones were recovered from the same levels. There were no axes, celts, or adzes. In the second level was found a small piece of hematite with a shallow groove cut into its surface. It was impossible to identify this fragment as any specific object. In the first level was found a rim fragment of a steatite bowl. The only stone objects with important cultural connations were atlatl weights. Atlatl weights. Five small atlatl weights were unearthed. One was found in each level from the first to the fifth. All of these weights were of the bar or rectangular type. None of the weights was whole. Each weight was represented by a fragment comprising approximately % to ^2 of its total length. The largest of these weights could have been no longer than M/i inches by 13A inches in width. The remaining four weights were shorter and narrower with the smallest one measuring about four inches in length by one inch in width. Fine-grained standstones were used for the manufacture of the bar weights. Except for the one found in the first level, none of the bars was carefully finished. They had been roughly pecked to shape and then rubbed to a semi-smooth finish. 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 185 Bar atlatl weights are commonly associated with Archaic manifestations in the Ohio Valley. Webb (1950, p. 551) considers the bar weight as the first type to be used in the Archaic shell-middens of Kentucky. Before the use of bar weights, the atlatl was probably made entirely of wood. BONE TOOLS Bone refuse was common in the first to the fourth levels of the excava- tion. Bone was not well preserved except in the first and second levels. Below the second level the deterioration of bone became progressively marked. The absence of bone below the fourth level may be due to the slight acidity of the soil and the long time span represented by those levels. Only 23 worked pieces of bone were found. In the first level were found eight specimens consisting of two small pins, three fragments of awls, one cut antler, and three antler tips. The second level produced eight objects consisting of four fragments of awls, three antler tips, and one worked antler base. In the third level were five specimens consisting of three frag- ments of awls, one notched pin, and one worked fragment of deer skull. The fourth level produced only one awl tip and one worked fragment unidenti- fiable as to form. All of the awls, with the exception of one turkey leg bone awl found in the third level, were made of splinters of deer bone. The tips were usually well formed but little attention had been given to the remainder of the tool. None of the bone tools was culturally distinctive. Table 4 gives the distribution of points, and Table 5 the distribution of all the artifacts found during the excavations at the Rohr Shelter. DISCUSSION The materials found in the various levels of the Rohr Shelter present some interesting insights into the cultural developments of the Upper Ohio Valley. Compressed within the first and second levels of the site were objects representative of three major cultural divisions— Early Woodland, Middle Woodland, and Late Prehistoric. The pottery sample from the excavation was not large, but distinctive types such as heavy grit-tempered sherds with cordmarked exteriors and interiors belonging to Early Woodland, lime- stone-tempered pottery of Middle Woodland times, and shell-tempered sherds common to Late Prehistoric manifestations were all represented. Stemmed projectile points with rounded bases were commonly associated with the heavy grit-tempered pottery. A variety of notched and stemmed points was in use during Middle Woodland times when limestone-tempered pottery was in vogue. Small triangular points found near the surface at Rohr Shelter belonged to the users of shell-tempered pottery during the Late Prehistoric period. The ceramic development represented in the Rohr Shelter covers a time span of nearly 2500 years. Pottery-using groups accounted for only about one-third of the debris deposit. Beginning at the third level and continuing through the seventh level were remains of non-pottery-using peoples who may have inhabited the shelter over a period nearly twice as long as that of the pottery makers. The distribution of projectile point types indicates that perhaps for a considerable time peoples of both the ceramic and non- TABLE 4. ROHR SHELTER. DISTRIBUTION OF POINTS BY LEVELS 186 Annals of Carnegie Museum vol. 35 £ oc ^ mo ® m f C !>■ mo CO i— ' h s oo H O0 c/0 CM ' X w c/o £ CM — < o ® CO 00 CM »■ C OO (_ GO (M CM CM 2 2 £ C 00 00 r c o H PQ *-* < CM 050 OJ be X i2 'ac W - ‘qj X £ X X X OOOCMOOrjiCMxO1-^ rH bl - CO ... C ^ gl ^ ° 00 mo CO CM —I CM OO O CO CO CM CM ^ CM t/3 as jj 8 G C S w cb — . Oh g j X y y .2 y S [3" J S ifO SO CM 00 I-H 1—1 O CM c/o pit ,—i CM SO 'bH y -1 I- •h’C CM b ^ 9 a c^Oh .2 qj 2 ^ o *H II Cb ( c/o w ^ J y cm t- co PQ S CO oo < c - h * y > ~ CM SO y Total 939 288 68 25 480 15 5 26 5 23 1172 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 187 ceramic traditions were using the shelter. It is entirely possible that there was little actual change in population; only the addition of new ideas concerning the making of pottery and new styles of points. The materials from the Rohr Shelter have given a number of insights into the development of Archaic culture in the Upper Ohio Valley. No other site now known in the area presents such a long temporal series. In the lowest levels of the site there was a predominant use of small side-notched and expanded-stem points. Occasionally a large, slender side- notched point was used. Small bifurcated-base points also appear to be part of the early Archaic complex. All of these point types, except those having the bifurcated base, continued to be used until after the introduction of pottery. The most apparent changes occurred in the late stages of the Archaic when large stemmed points became a common item of the complex. Such points were of both the parallel- and expanded-stem types. These types with modifications and refinements were carried over into the early pottery- using cultures. Small end-scrapers of the “thumbnail” type were important items in the tool kit throughout the Archaic, but they were absent from the major pottery levels. Side-scrapers were used from early Archaic through ceramic times. Drills were more abundant in levels containing pottery but they make their appearance midway in the non-pottery levels. Of particular interest are the atlatl weights found at the Rohr Shelter. All were of the bar type which has been considered an early form. None of the more elaborate winged types occurred. The Rohr Shelter undoubtedly presents only a limited picture of the total Archaic complex. In other areas of the eastern United States much richer Archaic complexes have been found. Unfortunately, in the Upper Ohio Valley no such sites have as yet been reported. As of this date it has been necessary to piece together the story of Archaic times from a number of small sites which have produced only a limited number of artifacts. Here no mention has been made of the relationship of the Rohr Shelter materials to those known from other sites in the Upper Ohio Valley and the eastern United States. In the following section of this article the Rohr materials have been correlated with those found at the Dixon Shelter and at other sites in the Upper Ohio Valley. RELATIONSHIP OF DIXON AND ROHR SEQUENCE IN THE UPPER OHIO VALLEY ARCHAIC In both the Dixon and Rohr shelters it was possible to trace a sequence of cultural events beginning in Archaic times and extending into the Late Prehistoric. Both shelters produced pottery sequences beginning with heavy grit-tempered cordmarked sherds which gave way to medium-thick grit- or limestone-tempered sherds which in turn disappeared with the introduction of shell-tempered pottery. It was, however, in the non-pottery producing levels that new information for understanding the cultural development in the Upper Ohio Valley was found. The sequence of stone artifacts, particu- 188 Annals of Carnegie Museum vol. 35 larly points and atlatl weights, has given new insights into the formation of the Archaic. The Archaic occupation at the Rohr Shelter seems to have begun at an earlier time than at the Dixon Shelter. At Rohr, the Archaic debris ac- counted for approximately one-half of the cultural deposit, while at Dixon it composed only about one-third of the total deposit. The Archaic picture at each shelter begins with the use of side-notched and expanded-stem points. Many of these points are very small. The presence of small bar atlatl weights at the Rohr Shelter indicates that throw- ing sticks were used to propel spears. These were probably spears with compound shafts like those used by the Eskimo in historic times. Consider- able conjecture has centered around the way in which these small Archaic points were used. Ritchie (1944, p. 321) believed that the many small points of the Archaic suggested the use of the bow. After working with the Shriver Site materials (Stewart and Dragoo, 1954, p. 110) which produced hun- dreds of small points but no atlatl weights among its many artifacts, I was inclined to agree with Ritchie on the early introduction of the bow. How- ever, the finding of bar weights at Rohr cast serious doubts upon this belief. In recent conversations with me, Mr. Ritchie expressed the view that the small points were used on compound spear shafts rather than on arrows as he had previously conjectured. Small points remained in use at both shelters well up into ceramic periods. Near the upper margin of the non-ceramic levels larger stemmed points be- came more prominent in the inventory of artifacts. Previous to the Dixon and Rohr excavations, large, crude stemmed points were generally con- sidered as being somewhat older than small notched and expanded-stem points. Their placement near the end of the Archaic at the Dixon and Rohr shelters calls for re-examination of some Archaic sites in the Upper Ohio Valley which were considered early because of the presence of such points. Another factor involved is the possible intermingling of two distinct cultural traditions. This point will be more clearly shown later in this section when other Archaic sites in the Upper Ohio Valley are discussed. At both the Dixon and Rohr shelters, small, finely worked end-scrapers were important items in the Archaic tool assemblage. They remained in use throughout the Archaic, but appeared in greater abundance during the early phases of the Archaic at Dixon and Rohr. Small end-scrapers have a long history beginning with Paleo-Indian cultures and continuing into Late Prehistoric times. Along with end-scrapers there were a number of less distinct types of scrapers found at Dixon and Rohr. Most noticeable among these are finely worked side-scrapers and small triangular cross-sectioned gravers or “burins.” Nearly every piece of flint found in the Dixon and Rohr excavations showed evidence of having been utilized for scraping purposes. A common feature of most Archaic sites in the Upper Ohio Valley is the great abundance of scrapers. It is my belief that such tools were used for extensive wood and bone working. If these perishable objects were available for study, the pic- ture of Archaic life would undoubtedly be enriched. 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 189 So far in this study data from only two stratified sites have been presented. The picture presented is of limited scope in understanding the Archaic in all its ramifications. Only small family groups and hunting parties occupied the shelters. The full cultural inventory of larger groups living in open villages was not present. Certain tools perhaps were not needed by the people during their stay in the shelters and are missing from our lists. The most important aspect of the shelter data is the chronological framework it provides to which we can attach specific artifact types. With such a frame- work it is possible to arrange other Archaic sites and materials in a logical sequence. Thus, new insights are obtained regarding the growth of specific items, and local developments and contacts can be more clearly traced. In the paragraphs which follow, attention will be focused upon the several Archaic manifestations known in the Upper Ohio Valley. (See Fig. 15.) Archaic sites in the southern Upper Ohio Valley Archaic materials in surface collections are known from several sites in the areas surrounding the Dixon and Rohr shelters, but only one of these sites is well enough known to present a picture of Archaic life. This site, known as the Gay Shriver site (36Grll), was discovered and worked by Dr. Paul R. Stewart of Waynesburg College. A preliminary report on the site was made by Stewart and Dragoo (1954). The Gay Shriver site is situated almost one mile southeast of Waynesburg in Greene County, Pa. It is situated on a flat-topped hill probably repre- senting a remnant of the so-called Harrisburg erosion surface, and commands a fine view of the surrounding area. It lies above one of the finest flowing springs in the county. The hill is so sandy that many years ago it was called the “Sand Hill”, and large quantities of sand were dug for use in road building. This sand comes from the weathering of sandstone which is located stratigraphically about half way from the base to the top of the Washington Formation and is of local significance only, and has been given no particular geological name. Several years ago the hill had a number of large productive chestnut trees upon it. On the slope of the hill still exists a small virgin forest of black and scarlet oak with some hickory. Forests of this type are generally called “Oak-Hickory Climaxes” (Illick, 1928). Before the extinction of chest- nuts, oak-hickory climaxes were often interspersed with chestnut trees. Such climaxes renew themselves for many thousand years and there is no reason to cloubt that such forests covered this hill during Archaic times. The forest floor has an association of blueberries ( Vaccinium ) and other acid-loving plants. Soil tests in the portion of the site covered by virgin forest showed strong acidity (pH4.5-5). The site proper covers less than an acre on the hilltop. Dr. Stewart over a period of many years searched for artifacts when portions of the site were under cultivation, and in 1951 he made a number of test pits. No artifacts were found more than two feet below the surface, and practically all speci- mens occur within plowshare depth. Approximately 1200 identifiable arti- facts plus several thousand crude pieces of flint were deposited in the museum at Waynesburg College. 190 Annals of Carnegie Museum vol. 35 Fig. 15. Major Upper Ohio Valley Archaic sites 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 191 The majority of the projectile points are of the corner-notched, corner- removed, or expanded-stem varieties. It is difficult to classify these points into well defined categories because of their crudeness and inconsistent shape. For example, a point may be corner-notched on one edge and side- notched on the other. Any one point could easily be classified in any one of the above mentioned varieties depending upon the inclination of the classifier. The largest points are no more than 2 inches long and most are less than 1^2 inches. It is not uncommon to find points less than one inch long. Many of these small points are quite blunt while others have well defined tips. Most are rather thick and nearly as wide as long. (See Fig. 16-17.) Among the minor point types the side-notched points are most numerous followed by those with stems. Pentagonal, bifurcated, and basal-notched points occur but are quite rare. There are only eight broad-stemmed and lanceolate points in the collection. (See Fig. 18-19.) A wide variety of material was used in the manufacture of the projectile points. In fact, it is notable that the chalcedony, chert, jasper, quartzite, etc., cover a far wider range of materials than those found on several Late Pre- historic sites located in the same general area. A number of the artifacts are made of poor quality dark chert found near Waynesburg. Outside of this local flint, no large number of artifacts is of any one type of material. A great variety of materials was noted in the Dixon and Rohr excavations and small surface collections from other Archaic sites manifest the same heterogeneity. Some drills and perforators are known from the Shriver site. Of the 16 unbroken drills, 10 have the notched base. Each of the remaining six drills has a straight to slightly expanded base. Both side and corner notches occur on the notched drills. Most of the drills are crudely fashioned. (See Fig. 20.) Scrapers and flake knives are abundant. Unspecialized chips and flakes were most commonly used in the manufacture of these scrapers and knives. Many of the flake knives resemble those used in Hopewellian times, but they are generally smaller and less finely made. Many of the scrapers show evidence of having been retouched several times. In the flint samples that have been gathered from the site, there are several hundred crude scrapers. Some of these show considerable use while others were apparently used only once and then tossed aside. The small end-scraper so common in the Archaic levels at Dixon and Rohr is not well represented in the Shriver collection. (See Fig. 20.) A number of small leaf-shaped blanks have been found at the Shriver site. The largest of these blanks are little more than two inches long while some are less than one inch long. Among the heavy polished stone tools are two full-grooved axes, two small grooved adzes, and three broken ax or adz fragments. All of these tools are badly battered and broken. Considering the number of specimens that have been found on the Shriver site axes and adzes were not an abundant tool. No axes or adzes were found in the Dixon and Rohr excavations. (See Fig. 21.) Pitted stones and hammerstones are common in the Shriver materials. The 192 Annals of Carnegie Museum vol. 35 Fig. 17. Shriver site. Typical small expanded-stem and notched points 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 193 Fig. 18. Shriver site. Stemmed and side-notched points Fig. 19. Shriver site. Blanks, wide-stemmed points, and points of weathered materials 194 Annals of Carnegie Museum vol. 35 Fig. 20. Shriver site. Scrapers and drills Fig. 21. Shriver site. Polished-stone tools 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 195 pitted stones differ from those found on many later sites in that they were made from angular blocks of sandstone rather than from smooth pebbles. Some of the hammerstones are crudely shaped spheres while others are only utilized pebbles. No pottery, bone implements, or bone refuse were found; and only a few tiny fragments of mussel shells were discovered. With the exception of five slate and three hematite fragments, no whole ornaments were found. Two of the slate fragments appear to be of Laurentian origin. One is a por- tion of a tube or possibly of an atlatl weight; the other could be a knife. The other three fragments are of local carbonaceous slate. The several excavations made at the Shriver site failed to uncover any true fire-pits or other sub-surface features. Scattered over almost the entire site, however, are stones showing evidence of firing. Burials are also unknown at the site. The high acidity of the soil has probably destroyed all human remains as well as all other bone refuse. With some conjecture we can reconstruct the village and portray the life of the inhabitants of the Shriver site. This site, although comparatively level, probably has been lowered a few inches by weathering, moving the artifacts downward a short distance. Chamberlin and Salisbury (1921, p. 84) estimated that the entire Mississippi Valley has been lowered approximately one-half foot in the past 5000 years. Granting that the top few inches of the site have disappeared since the Archaic occupation, so far no deep refuse pits, shelters, or burials have been found. Shelters were probably constructed of a light pole framework covered by bark or skins similar to the lean-to. Fires for cooking were built on the surface of the ground and banked by the stones which litter the site today. Refuse was not abundant and consisted mainly of bone which has been destroyed by the acid soil. Life was simple and re- volved around the gathering and hunting of food. There was little time for frills and the cultural elaboration which came in later times with the intro- duction of agriculture. There is a great possibility that the inhabitants did not live on this site during the winter season. The absence of evidence of substantial houses and deep fire-pits indicates that these people did not subject themselves to the cold winds which batter this hilltop during the winter. Very likely the inhabitants chose this locality for a summer and fall residence because it was near a fine spring and upon a hill of sour sand where underbrush was not rank. The chestnut trees along this and other ridges furnished them with considerable food in the fall. Wild pigeons and turkeys used such chestnut ridges as their roosting places and areas from which they gathered most of their provender. There was abundant small game as well as deer, elk, and bear to add to the menu. During the months of late summer and early fall there were various types of berries and nuts to spice their food with variety. With the beginning of the cool days of autumn, the people began their movement to more sheltered areas. It was during the winter months that rock shelters such as Dixon and Rohr saw their greatest occupa- tion. Remains of large and intense fires are a common feature of the rock shelter. 196 Annals of Carnegie Museum vol. 35 The material culture of the Shriver site shares many similarities with the materials found at the Dixon and Rohr shelters. This is particularly true of point types. The small notched and expanded-base points of Shriver are very much like those found in the lower and middle levels of the Dixon and Rohr Archaic levels. The broad-stemmed points considered as late in the Dixon and Rohr sequence are very rare at Shriver. On this basis I am now inclined to consider Shriver as representative of Middle Archaic times in the Upper Ohio Valley and not Late Archaic as I did in the earlier report on this site (Stewart and Dragoo, 1954, p. 110). The problem of the broad- stemmed point as a time marker will come up again later in this section with discussion of certain sites in the Ohio Valley proper where they have been considered as Early Archaic (Mayer-Oakes, 1955b, p. 19). The absence of atlatl weights at the Shriver site was mentioned in con- nection with the possible use of the bow in Archaic times. Considering the large number of artifacts found at Shriver over several years of collect- ing by Dr. Stewart, the only logical conclusions that can now be drawn are that either the weight and atlatl were not used or that the atlatl was constructed entirely of perishable material. I believe the latter to be more nearly correct. At the Rohr Shelter several crude bar weights were found throughout the middle and upper Archaic levels. Only in the lower levels were they missing. Since the bar weight has been considered as the earliest form of weight (Webb, 1950, p. 351), the absence of weights at Shriver may indicate an earlier Archaic manifestation than has been admitted in the past. Similiarities of the Shriver site materials with the Brewerton Focus of New York state (Ritchie, 1944) have been stated previously (Stewart and Dragoo, 1954). These similarities are particularly evident when the flint objects are compared. The thick, broad-bladed, notched and expanded-stem points; bifurcated base points; pentagonal points; notched-base drills; flake scrapers and knives; and small leaf-shaped blanks are nearly identical in form with those found at the Oberlander and Robinson sites by Ritchie (1944, p. 235-249). There are, however, important differences which will be discussed more thoroughly in the next section of this article. A number of small Archaic sites in southwestern Pennsylvania similar to Shriver were recorded by Carnegie Museum’s Upper Ohio Valley Archeo- logical Survey. None of these sites is well known. The materials in small surface collections from them are similar to those found at Shriver and in the Dixon and Rohr excavations. Some temporal differences may become evident when these sites are more thoroughly investigated and larger collec- tions of materials for study are available. Approximately twenty miles east of the Dixon Shelter in Garrett County, Md., a number of Archaic sites were reported by Frank R. Corliss, Jr. (1951, 1954) and recorded by Carnegie Museum. Garrett County is geo- graphically on the dividing line between the Ohio and Potomac watersheds. The major stream in the area is the Youghiogheny River of the Ohio drain- age. Most sites known for the area are located around the shores of Deep Creek Lake on the headwaters of the Youghiogheny. This lake, of artificial origin, has eroded into and removed the topsoil from nearly forty small sites along its shores. None of these sites has pro- 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 197 duced pottery. At 18Gal Corliss (1951, p. 1) found side- and corner-notched points made of local chert obtained from a quarry site about six miles away. Along with the points were found scrapers, drills, cache blades, hammer- stones, knives, and many flint flakes. An important feature of 18Gal was the finding of three cache deposits of artifacts. One of these caches contained 10 objects consisting of three corner-notched (expanded-stem) finished points, one turtle-back blank, and six roughly chipped large leaf-shaped blanks. The material for all these objects was a gray chert of local origin. When the Garrett County sites are considered as a group they present considerable variations in types of points. Represented in the collections are small expanded-stem, side-notched, and bifurcated points similar to those found at Dixon and Rohr. Some of the sites, particularly 18Ga31, pro- duced crude, tapered, stemmed points which are wide, with a thick cross- section. These points are not generally known from other Archaic mani- festations in the Upper Ohio Valley. At both Dixon and Rohr shelters a few points with wide, tapered stems were found in the upper levels of the ceramic zones. Thus, at Dixon and Rohr such points are late in the Archaic sequence and perhaps attest to contacts with the Garrett County area. Corliss (1954, p. 9) believed the large tapered-stem points to be representa- tive of Early Archaic. In light of the Dixon and Rohr evidence, combined with the finding of a winged bannerstone at 18Ga2 (Mayer-Oakes, 1955, p. 91) and the presence of cache blades, 1 am inclined to consider the materials as representative of the Late Archaic. In many respects the Gar- rett County materials are similar to those of Witthoft’s (1953) Transitional Period Cultures of eastern Pennsylvania and the coastal region. Broad points with contracting stem, and the presence of cache blades are common features of the Transitional Period Cultures. The broad, contracting-stem point tradition is generally absent from the Ohio Valley except on its ex- treme eastern fringes. The occurrence of such points in Garrett County, Md., and in the Dixon and Rohr sequence presents important evidence of the overlapping of two distinct cultural traditions, in eastern Pennsylvania and adjacent areas, the Transitional Period Cultures appear as a base upon which later Adena and Hopewell-like cultures developed in those areas. Except at the sites already mentioned, little is known about the Archaic in much of the Monongahela, Youghiogheny, and Cheat River drainages. Numerous small sites were located but materials from them are scarce. The few local collections checked by Carnegie Museum workers contain evidence of a wide distribution of the small notched and expancled-stem points found in the Archaic levels of the Dixon and Rohr shelters. Ohio Valley proper Archaic sites Our attention will now turn from the southern extensions of the Upper Ohio Valley to the Ohio River proper in western Pennsylvania, eastern Ohio, and northern West Virginia. Most of our information comes from three small shell-middens and several non-shell-bearing sites. Prior to 1953 almost the only well known Archaic site in the Upper Ohio Valley was the East Steubenville site (46Br31) on the east bank of the Ohio 198 Annals of Carnegie Museum vol. 35 River opposite Steubenville, Ohio (Mayer-Oakes, 1955a, p. 132). This site is a small shell-midden on a shelf projecting from a steep hillside more than a hundred feet above the present water-level and approximately 250 yards from the river. Excavation of the site was begun in 1938 by Walter Caldwell of Wellsburg, W. Va., and Wilbur Harris of Short Creek, W. Va., aided on several occasions by Walter Singer of Wellsburg and Elmer Fetzer of Weirton, W. Va. Notes taken by Fetzer and much of the material in the various collections were made available to Carnegie Museum for study. The most conspicuous feature of the site is the accumulation of refuse from the shellfish which probably constituted a major item of the people’s diet. The hunting of other animals was attested to by the finding of bones of deer, elk, bear, turkey, turtle, and fish. The only features other than the refuse accumulation, known for the site are a flexed human burial and two dog burials. The site appears to have been occupied only during Archaic times by a small band of people. The most numerous artifacts in the collections from East Steubenville were chipped-stone tools consisting mainly of projectile points. Two major types of projectile points from the site have been named by Mayer-Oakes (1955a, p. 140). One type, the “Steubenville Lanceolate”, is a crude lanceo- late form usually with a square base. The other type, “Steubenville Stemmed”, is a square-based point with a broad, parallel stem. The lanceolate type is not as well defined as is the stemmed type. Several of the lanceolate type (Mayer-Oakes, 1955a, p. 138) appeared to be weakly shouldered variations of the stemmed form. Most of the lanceolates are crude and have the general appearance of blanks rather than of finished tools. Those that are finished possess usage marks similar to those found on blades used as knives. The “Steubenville Stemmed” type is well defined in characteristics and seems to have a specific cultural and spatial distribution. (See Fig. 22.) A few small notched points as well as straight and expanded-base drills, scrapers, crude leaf-shaped blanks, and choppers are also part of the chipped- stone inventory. The notched points are similar to those found in the Archaic levels of Dixon and Rohr and at the Shriver site. (See Fig. 22.) Among the polished-stone tools found at East Steubenville are three- quarter-grooved adzes with pointed polls, plain adz, three-quarter-grooved ax, bi-pitted stones, hammerstones, and notched pebble net-sinkers. One crude bar-shaped stone may have been used as an atlatl weight. (See Fig. 23.) A number of bone items including various forms of awls, notched pins, conical projectile points, conical and cylindrical antler drifts, cut and per- forated bear jaw, perforated stemmed bone point (harpoon), and an un- usual split and perforated cylindrical awl were found. Almost all bone objects known for the Archaic in the LTpper Ohio Valley have come from shell refuse deposits where the alkalinity of the soil was maintained at a proper level by the presence of the shells. (See Fig. 23.) The second shell mound for which there is some information (Mayer- Oakes, 1955a, p. 132) is the New Cumberland site (46Hkl). It is a small shell-midden on top of the bluff overlooking the Ohio River just north of New Cumberland, W. Va. The site is now about 310 feet above the present river level and nearly one-half mile away. This site is known only from 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 199 lililslll iilii II \ j v i i i k & ii v. t I 4 □ 4 1 5 16 1 7 Fig. 22. East Steubenville site. “Steubenville Stemmed” and “Steubenville Lanceolate” points, notched points, drills, and blanks 200 Annals of Carnegie Museum vol. 35 Fig. 23. East Steubenville site. Bone and polished-stone tools 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 201 surface collections gathered by A. Michael, the land owner, and Elmer Fetzer of Weirton, W. Va. A tabulation was made of both collections and some of the items in the Fetzer collection were illustrated by Mayer-Oakes (1955a, p. 133). Of the 20 points tabulated from the New Cumberland site, 11 or 55.5% are of the broad-stemmed “Steubenville Stemmed” type. The remaining points are smaller notched and expanded-stem forms typical of the Archaic sites already discussed in this article. A straight drill, scrapers, and crude choppers of chipped stone are also known for the site. Included among the polished-stone tools of the New Cumberland site are one full-grooved ax, two crescent bannerstones, one bar atlatl weight, and two notched pebble net-sinkers. No bone implements, burials, or subsurface features are known for the site. The present disturbed condition of the site seems, to preclude the pos- sibility of future work adding materially to our knowledge of the site. The third shell-midden, and the only one for which there is available information derived from controlled excavations, is the Globe Hill site (46Hk34-l) excavated by Carnegie Museum in 1953 and reported upon by Mayer-Oakes (1955b). The Globe Hill site is on the east bank of the Ohio River opposite the north end of Blacks Island in Hancock County, W. Va. The site lies on a hilltop nearly 150 feet above the present river bank. The main feature of the site is. a shell-midden about 80 feet in length and 50 feet in width. The deepest deposit of shell is only about 20 inches near the south end of the midden. The entire shell deposit thins out toward the margins. Mixed in with the shell are various artifacts of chipped and polished stone, heat- cracked rocks, and bone refuse. Some artifacts have been scattered on the surface surrounding the midden but there is no great concentration of such items. Although the site was carefully excavated by levels, no important differ- ence in the artifacts was noted from the top to the bottom levels. No features such as pits or burials were found during the excavation in 1953, but frag- ments of human bone have been found recently by amateurs digging in an area not worked at that time. The site appears to represent a single occupa- tion by a small band of people who depended basically upon the shellfish for subsistence. Chipped-stone objects were abundant in the shell debris at Globe Hill. The most common objects were projectile points. Broad-stemmed points of the “Steubenville Stemmed” type made up 15% of the sample and lanceolate points of the “Steubenville Lanceolate” type accounted for 14% (Mayer- Oakes, 1955b, p. 21.). The remaining 71% of the points consisted of 38% corner-notched or expanded-stem points, 15% side-notched, 1% single- shouldered stemmed points, 1% triangular points, and 16% crude lanceo- late points. Again, as was noted for the East Steubenville and New Cumber- land sites, the lanceolate points are not a well defined category. Some of the points classified as lanceolate by Mayer-Oakes (1955b, p. 9) appear to be weakly shouldered stemmed points while those considered as crude lanceo- 202 Annals of Carnegie Museum vol. 35 late are more like blanks than true projectile points. The ‘‘Steubenville Stemmed”, however, seems to be a well defined type. (See Fig. 24.) The corner-notched or expanded-stem points and the side-notched points are similar to those recorded for Dixon and Rohr and the Shriver site. (See Fig. 24.) A number of scrapers and a fewT drills were found at Globe Hill. The most common form was the side-scraper usually of the flake variety. Three spoke-shave scrapers and several thin flake knives were also recorded. No end-scrapers were noted at the site. Only one expanded-base and one straight- base drill were found during the excavation, but one slightly expanded-base drill from the site is part of the Elmer Fetzer collection. A number of interesting polished-stone items are known for the Globe Hill site. These include plain adz, beveled adz, grooved adz, celt, crescent bannerstone, bar atlatl weight, notched pebble net-sinker, stone disk, bi- pitted stone, hammerstone, polished hematite concretion, sandstone shaft smoother, worked cannel coal, slate reel-shaped gorget fragment, and small hand-pestle. A three-quarter-grooved ax was found about one hundred yards from the midden deposit. This item may not belong to the complex. The grooved adzes are crude examples of the type. They do not have well de- fined raised ridges along the groove. The bar weights are rectangular pieces of crudely smoothed stone. The crescent bannerstone is the most distinctive spearthrower weight. (See Fig. 25.) Among 3376 bone fragments found at Globe Hill there are a few worked pieces including such items as a cylindrical drift, antler flakes, awls, spatula, perforated deer astragalus and cut dog jaw. The most interesting item is the perforated deer astragalus which is similar to items known in Lamoka in New York (Ritchie, 1944) and some Tennessee Valley sites (Kneberg and Lewis, 1955). Worth mention, also, is the presence of dogs. Globe Hill and East Steubenville are the only two Archaic sites in the Upper Ohio Valley for which dogs have been reported. (See Fig. 25.) A few small shell-midden deposits are known for the hilltop areas sur- rounding the Globe Hill site. No materials are known for these sites. Only the three shell-middens just discussed have produced enough information to make comparative studies possible. On the basis of these three shell- middens and selected materials from other sites Mayer-Oakes (1955a, 1955b) formulated a complex which he called the “Panhandle Archaic.” He be- lieved the “Steubenville Lanceolate” and “Steubenville Stemmed” points were the end product of the Paleo-lndian period and the earliest Archaic points in the Upper Ohio Valley. These points were derived from late Paleo-lndian “Scottsbluff” and “Starved Rock Lanceolate” types, which came into the area from the west at about the time of the post-glacial climatic maximum. Thus, to Mayer-Oakes, the shell-midden sites represent the adoption of new food resources based on the collecting of shellfish and the lessening of hunting activities formerly predominant in the Paleo-lndian period. Since stemmed and lanceolate points represent the bulk of the known col- lection for East Steubenville, Mayer-Oakes (1955b, p. 20) considered the 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 203 ft* «• « : il.ai mi : I . ****4#*4#46 #♦> * m *14111 4 § H 'm t -j. ^ /■ Pi # 4* fSl m MifiP Rl 1 2 ZJ- ■ L L_ 8 I? |l0 | Fig. 24. Globe Hill site. “Steubenville Stemmed” and “Steubenville Lanceolate” points, blanks, and notched points 204 Annals of Carnegie Museum vol. 35 Fig. 25. Globe Hill site. Bone ancl polished-stone tools 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 205 East Steubenville site as older than Globe Hill where notched and expanded- stem points made up a greater proportion of the sample. On this basis the Shriver site was considered as quite late, as only a few stemmed and lanceolate points were present. To support his chronological sequence based on point types, Mayer-Oakes cited the presence of the grooved adz as the major ground-stone tool at East Steubenville and equated it with the presence in Lamoka of the beveled adz as the primary tool there. He assumed that western New York Lamoka was the origin of the beveled adz in Early Archaic times; therefore, the presence of a beveled adz at Globe Hill and not at East Steubenville became support for the later relative date for Globe Hill. The presence of the crescent-shapecl bannerstone at Globe Hill and not at East Steubenville was cited as additional evidence for a later date. There are several elements in the “Panhandle Archaic” complex which tend to confuse the picture both culturally and chronologically. The “Steu- benville Stemmed” and “Steubenville Lanceolate” points are distinctive for the “Panhandle Complex.” They occur in only minor numbers or not at all at other Archaic sites. If such points are considered as the end product of the Paleo-Indian period, we must assume that they remained in use for a period covering several thousands of years during the Archaic. The Globe Hill site which produced several of these points also possesses traits, such as the crescent-shaped bannerstone, which are generally considered as late. Both Globe Hill and East Steubenville possess a well developed ground- stone complex. If we accept the assumption that ground-stone and polished- stone tools are not common until late Archaic times, then we are faced with the problem of a supposedly old point complex associated with much later polished-stone items. The small notched and expanded-stem points found at Globe Hill have been considered as an indication of lateness (Mayer-Oakes, 1955b). The Dixon and Rohr sequence, however, presents evidence that these small points were in use in the Upper Ohio Valley for a very long period of time and that their presence at a site does not necessarily indicate lateness. The Dixon and Rohr sequence is of little aid in the chronological placement of “Steubenville Stemmed” and “Steubenville Lanceolate” types. No typical points of these types were found. All of the wide-stemmed points which could be remotely associated with such types were confined to the upper Archaic and lower ceramic levels. The presence of beveled adzes at Globe Hill and not at East Steubenville has entered into the picture as a factor in the chronological ordering of the two sites. Mayer-Oakes (1955b, p. 22) has assumed that the East Steubenville site is earlier because it does not have the beveled adzes. In the Lamoka and Laurentian cultures of New York the beveled adz was found in early levels and continued in use over a considerable span in time. Since the beveled adz is known to be early in Archaic contexts, we must conjecture that the grooved adz which occurs at East Steubenville is earlier than the beveled adz. Thus, the temporal, spatial, and cultural relationships of the grooved adz become of crucial importance in the determination of the time and cul- 206 Annals of Carnegie Museum vol. 35 tural affiliations of the “Steubenville Stemmed” and “Steubenville Lanceo- late” points. The grooved adz is best known from sites in New England, New Jersey, and a few scattered sites in eastern New York and eastern Pennsylvania. The grooved adz was present in the late Archaic Coens-Crispin complex of New Jersey. (Cross, 1941). In eastern Pennsylvania grooved adzes have been found in the Philadelphia area and the Schuylkill Valley where they are perhaps related to the late Coens-Crispin complex of New Jersey (Witthoft, 1957, personal communication). According to Rouse (1957, personal com- munication) the grooved adz is common in collections from Connecticut, and perhaps belongs to the Sebonac Focus of Late Woodland. In eastern New York this tool is more gouge -like than adz-like, having a more or less concave lip and trough-shaped surface. The grooved gouge occurs sparingly in the Hudson Valley and in the Lake Champlain area. Ritchie (1957, per- sonal communication) relates this item to the Vosburg complex which he believes to extend from middle to late Archaic. The grooved adz is an unknown tool in the Archaic of the lower Ohio Valley and the Tennessee Valley. Thus, its development took place in the Northeast, and from available evidence this occurred in late Archaic times. Whether it developed first in the Upper Ohio Valley or in New England is as yet unanswerable. Archaic materials similar to those known for East Steubenville and Globe Hill have been found at several sites along the Ohio River. There are only four sites, however, for which sufficient information is available to make comparisons. These sites are the Half-Moon site (46Br29), Watson site (46Hk34), McKees Rocks (36A1 16), and the Wheeling College site (460h22). No definite shell-midden deposits are known at these sites, but they all produce Archaic materials as well as items from later manifestations. The Half-Moon site (46Br29) is an extensive bottom-land site in Brooke County, W. Va., about four miles southwest of Weirton (Fetzer and Mayer- Oakes, 1951. Mayer-Oakes, 1955). The site once consisted of four small Adena burial mounds and several areas containing culture debris. Recent construction work has destroyed much of the site including all but one of the mounds. Most of the materials from this site which can be assigned to the Archaic on a typological basis were found on the surface by Elmer Fetzer of Weirton. Included among these objects are the crescent-shaped bannerstone, three- quarter grooved adz, full-grooved ax, “Steubenville Stemmed” and “Steuben- ville Lanceolate” points, net-sinkers, and pitted stones. All of these materials are similar to those found at the East Steubenville and Globe Hill shell- middens. The Half-Moon site, however, is best known for its Adena mounds and associated objects such as reel-shaped gorgets, tubular pipes, celts, hematite hemispheres, copper beads, and Half-Moon ware (Fayette Thick) pottery. The McKees Rocks site (36A1 16) is similar to the Half-Moon site. This site is on the west side of the Ohio River at the mouth of Chartiers Creek approximately three miles northwest of Pittsburgh. It is a large site situated on high ground consisting of an Adena mound and areas producing village 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 207 debris of several time periods. The mound was partially excavated by Car- negie Museum in 1896 and the artifacts have been made available to several workers for study. The first major report on the site was made by Carpenter (1951) and two subsequent analyses have been made by Mayer-Oakes (1955a) and M (Michael (1956). Included among the many artifacts excavated in 1896 at the McKees Rocks mound were several “Steubenville Stemmed,” “Steubenville Lanceolate,” and small - notched and expanded-stem points which typologically can be assigned to the Archaic. Also present are the grooved adz and the beveled adz. The presence of this material in the mound fill has been considered as due to the scraping of a nearby Archaic village for earth (Mayer-Oakes, 1955a, p. 152). Many of the objects from the mound are of Adena types and include reel-shaped gorgets, hematite cones, celts, shell beads, bone awls, and stemmed points of typical Adena form. Thus, in the McKees Rocks mound materials, there were representative items from three known com- plexes—Panhandle Archaic, Laurentian, and Adena. A fourth complex, repre- sented in surface collections from the site and one intrusive burial in the top of the mound, is the Late Prehistoric. This last mentioned material, however, is quite late in the occupation of McKees Rocks. (See Fig. 26-27.) On the terrace along the Ohio River just north of the Globe Hill site in Hancock County, W. Va., is the large Watson site (Dragoo, 1956). The main features of this site belong to the Middle Woodland period and include at least two stone mounds, one of which was excavated by Carnegie Museum in 1953. In the lowest levels of this excavation were found several pottery sherds of Half-Moon ware which is representative of Early Woodland. Stemmed points of typical Adena form and a few “Steubenville Stemmed,” and “Steubenville Lanceolate” points have been found on the surface along with a fragment of a grooved adz. A large number of points of the notched and expanded-stem types also have been collected from the surface. Al- though the Archaic elements of the Watson site are only superficially known, there is a similarity in the contents and the sequence (from Panhandle Archaic to Early Woodland) with the Half-Moon and McKees Rocks sites. Several sites in Beaver County, Pa., lying in the Ohio Valley between the Watson site and the McKees Rocks site have produced Archaic materials. None of these sites, however, is well known and they are represented only by surface materials. Site 36Bv22 has produced two grooved adzes, a beveled adz, five plain adzes, a crescent bannerstone, broad-stemmed points, and notched points. At site 36Bv52 a grooved adz and several “Steubenville Stemmed” and “Steubenville Lanceolate” points have been found (Mayer- Oakes, 1955a, p. 141). From site 36Bv3Q, currently under investigation by Carnegie Museum, a wide range of Archaic and Early Woodland materials has been collected. Among these materials are grooved adzes, full-grooved axes, celts, crude choppers, many scrapers, including burin-like tools, broad- stemmed and lanceolate points, notched points, crescent bannerstones, and Early Woodland pottery. Future study of these sites should add much to our knowledge of the Archaic. In marked contrast to the sites discussed above is the Wheeling College site (460h22) at Wheeling, W. Va., only a few miles down the Ohio River 208 Annals of Carnegie Museum vol. 35 Fig. 26. McKees Rocks Mound. Stemmed and lanceolate points and blanks Fig. 27. McKees Rocks Mound. Polished-stone tools 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 209 from the East Steubenville site. The debris of this site covers an extensive area of a hilltop overlooking Wheeling Creek. No mounds or shell-middens are known on or near the site. Over a period of many years several local collectors gathered artifacts from the area. A recent report by Clifford M. Lewis (1955) gives the contents of these collections and other pertinent data concerning the site. The range of materials known for the Wheeling College site indicates that more than one culture was represented, but the great bulk of the materials consists of small notched and expanded-stem points similar to those found in the Dixon and Rohr excavations and common on the Greene County, Pa., sites such as Shriver. Only a few points could possibly be considered as examples of the “Steubenville Stemmed” or “Steubenville Lanceolate” types. Missing from the list of items from the Wheeling College site are the grooved adz and the crescent bannerstone common to the sites of the “Pan- handle Archaic.” Among the ground- and polished-stone tools from the site are full-grooved axes, crude choppers, hammerstones, pitted stones, notched pebble net-sinker, and smoothed pieces of hematite. Other items including the three-quarter-grooved axes, celts, pottery sherds, and gorget fragments have been found on portions of the site, but such items do not seem to be associated with the notched and expanded-stem points from the hilltop area of the site. Lewis (1955, p. 7) noted that pottery and artifacts generally believed to be later in time than the Archaic were found on the bottom-land. On the basis of present evidence, the Archaic materials from the Wheeling College site are more closely related to the lower Archaic levels of Dixon and Rohr and to sites in Greene County, Pa., than to the “Panhandle Archaic” as represented at East Steubenville and Globe Hill. Thus, the materials from the Wheeling College site can be related to Archaic mani- festations found over most of the Upper Ohio Valley and shown to be old in the Dixon and Rohr sequence. Returning now to the overall picture of the Archaic in the Upper Ohio Valley proper, we are faced with the problem of two major Archaic com- plexes (Laurentian and Panhandle Archaic) inhabiting the same area and both supposedly having considerable antiquity. The Laurentian culture, as manifested at the Shriver site, Wheeling College site, and Dixon and Rohr shelters, has a wide distribution throughout the Upper Ohio Valley and can be closely correlated with the Laurentian of New York state where radiocarbon dates have established its presence there at around 3500 B.C. (Ritchie, 1955). The small notched and expanded-stem points so common to Laurentian culture were found beginning at the lowest levels in the Dixon and Rohr sequence where they were also asso- ciated with the simple bar atlatl weight. The full-grooved ax is the common polished-stone tool present at the southern Laurentian sites while the plain or beveled adz is the prevalent tool on the northern Upper Ohio Valley and New York sites. The grooved adz is not common and when present is con- fined to sites near the “Panhandle Archaic” sites. Also associated with the 210 Annals of Carnegie Museum vol. 35 grooved adz are stemmed points similar to the “Steubenville Stemmed.” Such points in the Dixon and Rohr sequence occur in the late Archaic levels and early ceramic levels. The position of Laurentian materials in the lowest levels of stratified deposits, the distribution of such materials over much of the Ohio Valley and Northeast, the simple form of certain objects such as the bar atlatl weight at some sites, and the presence of local developments within the general Laurentian framework, all indicate that the Laurentian was an old and well established culture which in its late periods came into contact with new cultural traditions. Such contracts account for the variable artifact inventories present at some widely scattered Laurentian sites. Considerable antiquity also has been claimed for the “Panhandle Archaic” by Mayer-Oakes (1955b, p. 19) who has considered the “Steubenville Stemmed” and “Steubenville Lanceolate” points as the end product of Paleo-Indian culture surviving into Archaic times. On the basis of the in- formation presented in preceding paragraphs of this article, l am unable to see such antiquity for the “Panhandle Archaic.” My reasons for this belief can be summarized as follows: 1. The “Panhandle Archaic” can be related to Paleo-Indian culture only by a typological resemblance of the “Steuben- ville Stemmed” and “Steubenville Lanceolate” points to similar types such as “Scottsbluff” in the Plains. 2. Similar points were found in the upper levels of the Dixon and Rohr shelters where they were considered as transitional forms from Late Archaic to Early Woodland. 3. Typologically the “Steubenville Stemmed” and “Steubenville Lanceolate” points also can be considered as variations of stemmed points and lanceolate points which occur at late shell-middens in Kentucky (Webb, 1946) and on Early Wood- land sites in the Upper Ohio Valley which appear to be representative of Adena. 4. The distribution of “Panhandle Archaic” sites is confined to or is in close proximity to the Ohio Valley proper and closely parallels the distribution of Early Woodland or more specifically, Adena sites. 5. The presence of the crescent bannerstone and grooved adz as typical traits of the “Panhandle Archaic” also indicates lateness for the complex. At a site currently under investigation by Dr. Paul R. Stewart of Waynes- burg College, additional evidence for the lateness of the “Panhandle Archaic” is being discovered. On this site, known as the Sproat site (36Grl6), in Greene County, Pa., several “Steubenville Stemmed” and “Steubenville Lanceolate” points have been found in association with fragments of tube pipes and gorgets. Some of the stemmed points are similar to those found on Adena sites. A few notched and expanded-stem points were found at the site but they are in the minority. Future work at this site may do much to clarify the chronological position of the “Panhandle Archaic.” The oldest and most widespread Archaic culture in the Upper Ohio Valley proper was the Laurentian. The strategic position of sites in the Ohio Valley near the confluence of several major tributaries entering the Ohio River from all directions permitted contacts with distant areas. These contacts led to regional specializations, particularly in polished-stone tools. For example, grooved axes are most common on southern sites while the 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 211 plain and beveled adzes are typical of the northern sites. The grooved adz and associated stemmed points represent a new cultural thrust into the Upper Ohio Valley from probably the northeast in Late Archaic times. This new culture blended with the resident Laurentian to form the “Panhandle Archaic” which became the base upon which the Early Woodland developed. ARCHAIC SITES IN THE NORTHERN UPPER OHIO VALLEY The northern Upper Ohio Valley is drained by two major rivers and their tributaries. The eastern and northernmost portions of the area are drained by the Allegheny River and its major tributaries— the Kiskiminitas and Conemaugh rivers, the Clarion River, and French Creek. The western area is drained by the Beaver River and its main tributaries, the Shenango and Mahoning rivers. Prior to Carnegie Museum’s Upper Ohio Valley Archeological Survey program, started in 1950, little was known about the cultures represented in the northern portion of the Upper Ohio Valley. Large areas still remain to be investigated. Most of the information pertaining to the Archaic was obtained from the surface of sites by the above survey and by recording and photographing private collections. No major excavation has been con- ducted on any Archaic site in the area. Butler (1936, p. 27) recorded the presence of materials of a hunting and fishing people belonging to the Archaic in Erie County, Pa. Her informa- tion was based on the presence of beveled adzes and “narrow-bladed” points in collections of surface finds. Carpenter (1942) reported on the results of a site survey in northwestern Pennsylvania. He indicated the presence of several small sites in which Laurentian materials had been found. Con- sidering the beveled adz as the main Archaic trait. Carpenter (1942, p. 23) stated: “Unfortunately, all site collections that yielded beveled adzes were so intermixed with the implements of other cultures that the former artifact represented the sole trait for the Archaic Pattern. Beveled adzes have been found in collections from Tionesta, West Hickory, Sugar Run, Tracy Run, Hodge Run, and Kiantone. At no one place were they plentiful, and the largest number from any site was three. The characteristic choppers and antler-tine pendants were absent, but the thin celt-like blade or scraper characteristic of Archaic is found associated with, and greater numbers than, the beveled adze.” Carnegie Museum’s survey in the Allegheny drainage produced similar information concerning the presence of narrow-bladed stemmed and notched points, plain and beveled adzes, and crude chopping tools (Mayer-Oakes, 1955a, p. 50). Sites 8 and 9 in Chautauqua County, N. Y., and sites 1, 2, 3, 7 and 9 in Forest County, Pa., contained a high proportion of these artifacts. In the Conemaugh drainage, sites 36In7, 36In8, 36In9; 36Cb2 and 36Cb3 pro- duced similar materials. Most of the information about the Archaic in the Allegheny Valley comes from the Siggins site (36Fol) and the County Home site (36Fo3) reported by Mayer-Oakes (1955a, p. 52). Both are large sites on high terraces along the Allegheny River. Much of the material from both sites appears to be of Archaic origin, but some items represent later Woodland occupations. 212 Annals of Carnegie Museum vol. 35 Typical Archaic objects from the Siggins site (36Fol) are beveled adzes, plain adzes, crude choppers, full-grooved axes, bi-pitted stones, hammer- stones, bannerstone fragments, notched pebbles, stemmed points, notched points, small bifurcated points, straight and expanded-base drills, end- scrapers, and side-scrapers. Also included in the collection from this site are several tube pipe fragments, elbow pipes, gorget and pendant frag- ments, celts, and steatite sherds. Most of the notched and expanded-stem points are similar to Laurentian, but the parallel-stemmed points are more like Early Woodland forms and are probably associated with the pipe and gorget fragments. On the County Home site (36Fo3) have been found such items as beveled adzes, plain adzes, full-grooved axes, notched and grooved pebbles, hammer- stones, winged bannerstones, notched and stemmed points, many end- scrapers, and several steatite sherds. Thus, this site presents a mixture of Archaic and Early Woodland traits. The presence of steatite vessel fragments at 36Fol and 36Fo3 is suggestive of the Transitional Culture of eastern Pennsylvania described by Witthoft (1953). Missing are the wide-stemmed points so common to that complex. Because of this factor, Mayer-Oakes (1955a, p. 53) hesitated to postulate a transitional period for the above sites, but considered the use of steatite vessels to be part of the general late Archaic complex. Due to the close proximity of these sites to the Susquehanna drainage where the Transi- tional Culture is present, contacts between the Upper Ohio Valley and the East in late Archaic times seems most likely. A similar situation was previously mentioned in this article for the materials from Garrett County, Md., and the presence of broad-stemmed points and sherds of steatite vessels in the transitional (Archaic and Early Woodland zones) of the Dixon and Rohr sequence. The Boyer’s Run Shelter (36Ve6) in Venango County, Pa., has produced a few Archaic artifacts at its lowest levels (Thompson, 1955). Although most of the items found in the excavation of this shelter by A. C. Thompson be- longed to ceramic periods, a few Laurentian-like notched and expanded- stem points are represented in the collection. The Allegheny Valley Archaic materials are basically similar to those of the Laurentian in New York. Unfortunately, only materials from late sites are represented in the collections that have been studied. Until further work is done in the area, it is impossible to establish a valid Archaic sequence. In the Beaver Valley many small sites located on benches and terraces along both major and minor streams and producing Archaic materials have been found. Only a few of these sites have been investigated and none has been excavated. During Carnegie Museum’s Upper Ohio Valley Archeological Survey, several Archaic sites and local collections of artifacts from them were re- corded for the Shenango Reservoir area in Trumbull County, Ohio, and Mercer County, Pa. (Mayer-Oakes, 1953). On five sites (33Tr4, 36Mel6, 36Mel7, 36Me26, and 36Me28) were found both Archaic and Early Wood- land materials. On four of these sites Middle Woodland items were also 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 215 present. Thus, our knowledge of these cultures, confined to sorting of col- lections on a typological basis, is not clear because of this mixing of artifacts. Included among the Archaic artifacts from the above sites are beveled adzes, plain adzes, full-groovecl axes, several types of bannerstones (winged, pick-shaped, rectangular with oval cross-section, and crescent), bi-pitted stones, hammers tones, end-scrapers, drills, bifurcated-base points, stemmed points, and notched points. The bifurcated-base and notched points are similar to those found in the Archaic levels of the Dixon and Rohr sequence and other Laurentian-like sites previously mentioned in this report. The stemmed points are like those from Early Woodland sites and are probably asso- ciated with the gorgets, pendants, fire-clay tubes, and cache blades also found on these sites. Other important sites in the Beaver Valley were reported by Mayer-Oakes (1955a). Site 33Mh4 in Mahoning County, Ohio appears to have been occu- pied both in Archaic and Early Woodland times or was perhaps a transi- tional site. Mayer-Oakes (1955a, p. 74) considered the plain and beveled adz, bannerstones, scrapers, and “Steubenville Stemmed” points found at this site as typical Archaic artifacts. In his summary of Beaver Valley sites Mayer-Oakes (1955a, p. 78) stated: ‘The overall impression one gets for this area is of an abundance of small camp sites on benches or terraces along both major and minor streams. Most of these sites represent very brief and probably intermittent occupations by wander- ing hunters. Surface materials are of chipped and ground stone with the lanceolate projectile points representing the continuation of late Paleo-Indian styles into early Archaic times. The fuller complex of adzes, axes, bannerstones, and “Steuben- ville Stemmed” points indicates an early period within the Archaic. A later Laurentian-like unit possibly represents a developed Archaic period.” Again, as was the common occurrence in the Ohio Valley proper, it is important to note that the “Steubenville Stemmed” points are found on sites with late Archaic and Early Woodland items. Thus, their associations are with transitional sites late in Archaic times and not with materials that reflect connections with the Paleo-Indian period. The grooved adz, which occurs in association with the “Steubenville Stemmed” points in the Ohio Valley proper, is a rare item in the Beaver Valley; but it has been recorded for site 36Me9 where stemmed points were also present. The plain adz, beveled adz, and full-grooved ax are common items from the Beaver Valley sites. At a number of small sites now under investigation by Carnegie Museum in the Beaver Valley, typical Laurentian notched, bifurcated and expanded- stem points occur in association with a simple ground stone complex con- sisting of plain adzes, grooved axes, and hammerstones. Missing from many sites are the more elaborate forms of bannerstones and the wide stemmed points. The varied inventories from these sites indicate that they represent occupations scattered over a long time during the Archaic. SUMMARY The Archaic complexes of the Upper Ohio Valley may be summarized as follows: E The earliest and longest Archaic occupation of the Upper Ohio 214 Annals of Carnegie Museum vol. 35 Valley was by peoples possessing a Laurentian-like culture. Their subsist- ence was based upon a hunting, fishing, and gathering economy. Most groups depended upon game from the forests, but some peoples living in the Ohio Valley proper also collected shellfish. The smallness of the villages indicates that during most of the Archaic the population was never large. Typical manifestations of the early Archaic are found in the lower levels of the Dixon and Rohr shelters and at numerous small surface sites scattered over the entire Upper Ohio Valley. Basic items in their tool kits included small notched, bifurcated, and expanded-stem points, a variety of scrapers in- cluding end-scrapers, full-grooved axes, plain adzes, bar atlatl weights, bi- pitted stones, and hammerstones. In late Archaic times new items such as the beveled adz and elaborate forms of the bannerstone, such as the crescent and winged types, were added to the basic items. Many Laurentian-like Archaic groups may have persisted in the more isolated areas of the Upper Ohio Valley well into Early Woodland times. 2. During late Archaic times along the eastern and southern portions of the Ohio Valley in close contact with the Potomac and Susquehanna River drainages, influences, and perhaps actual groups of people, of the Transi- tional Cultures of eastern Pennsylvania and the seaboard introduced to the Archaic new items including broad, tapered-stemmed points, cache blades, and steatite vessels. 3. In the Upper Ohio Valley proper at the late Archaic level, new peoples using broad-stemmed points, lanceolate blades, grooved adzes, and crescent bannerstones mingled with resident Laurentian-like groups to form the “Panhandle Archaic,” which in turn appears to be the base upon which Early Woodland (particularly Adena) developed in the area. The distribution of objects and sites associated with the “Panhandle Archaic” closely follows the distribution of Early Woodland, and materials of both complexes are commonly found on the same sites. Although the major occupation of the “Panhandle Archaic” occurs in the Ohio Valley proper, several sites in the Beaver River drainage show a mixture of this complex with Early Woodland materials. 4. A number of local developments and specializations of certain tools occurred in the various areas of the Upper Ohio Valley during the Archaic. On sites in the northern Upper Ohio Valley the common polished-stone tool is the plain and beveled adz which is the most common Archaic tool in the adjoining area of New York. Adzes occur on the southern sites but are less common than the full-grooved ax which is widely distributed on sites down the Ohio Valley and common on the shell-middens in Ken- tucky. The full-grooved ax is probably the oldest polished-stone, chop- ping tool. The grooved adz was used mainly in the Ohio Valley proper during late Archaic times. Its origin and relationships to other areas are not yet clear. 5. No positive evidence can be shown for direct contact between the Archaic and the earlier Paleo-Indian cultures. This may be due either to inadequate sampling, or a true chasm actually may have existed between 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 215 the two periods in the Upper Ohio Valley. Only continued and intensive research can solve this problem. Having set forth the major tenets of the Archaic in the Upper Ohio Valley, attention must be turned now to the relationship of this material to surrounding areas and in general to the Archaic as known in eastern North America. ARCHAIC MANIFESTATIONS OF EASTERN NORTH AMERICA The Upper Ohio Valley has been the path for movements of many peoples over long periods of time. It is the natural gateway and most direct route to the northeast from the Mississippi and Ohio Valleys. Thus, Archaic assem- blages of the Upper Ohio Valley have important relationships with mani- festations in surrounding areas in eastern North America. In the following paragraphs these relationships and their implications for understanding the origin and spread of the Archaic are investigated. (See Fig. 28.) The closest area to the Upper Ohio Valley for which detailed comparative information exists on the Archaic is New York. The recognition and elucida- tion of several diversified Archaic manifestations in New York developed through the excavations and publications of William A. Ritchie under the auspices of the Rochester Museum of Arts and Sciences between 1925 and 1940, and since then in his capacity as New York State Archeologist. Ritchie (1955) defined an Archaic period culture sequence involving two distinctive traditions which he named Lamoka and Laurentian. While Lamoka was given priority in time, its later interaction with a phase of Laurentian was believed to have occurred at Frontenac Island. Ritchie (1955, p. 3) envisioned the Frontenac complex as a product of an acculturational situation where various recombinations of older items and new traits occurred. The Lamoka tradition (Ritchie, 1932, 1955) was characterized by a primary concentration on a bone and antler industry. Rough and polished stone tools were moderately emphasized, while chipped-stone tools received only weak expression. Most Lamoka bone work was of a utilitarian nature relating to hunting and fishing activities. A few items, however, were ornamental forms, the most outstanding being a unique and diagnostic class of pendant-like ob- jects often decorated with red ocher (Ritchie, 1932, p. 106). Fishing equip- ment of bone included both hooks and gorges. There were bone awls of various styles but those made from deer scapula, and turkey radius were particularly numerous and distinctive. Other objects of bone included antler punches, beaver incisor chisels, bone whistles, and perforated deer astragali. The Lamoka polished-stone inventory included crudely executed celts, thin celt-like scrapers, plano-convex adzes, and beveled adzes. This last item Ritchie (1955, p. 5) considered as diagnostic of Lamoka and as having originated from the plano-convex adz in western New York where it has been found in large numbers. Other rough stone items in Lamoka included hammerstones, bi-pitted stones, anvil stones, crude choppers, notched net-sinkers, multi-pitted stones, mu Hers, cylindrical pestles, flat grinding stones, and basin-shaped and trough-shaped mortars (Ritchie, 1955, p. 5). These last items coupled with 216 Annals of Carnegie Museum vol. 35 Fig. 28. Major central and northeast Archaic manifestation 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 217 the finding of acorns and other plant remains demonstrated the importance of wild plants in Lamoka economy. Ritchie (1932, 1955) stated that the typical Lamoka projectile point was crudely percussion-chipped, small, weakly side-notched or stemmed, slender- bladed and with a thick base. Such points were used on spears rather than on arrows. According to Ritchie (1955, p. 5) ceramics, copper or shell objects, har- poons, barbed bone points, bannerstones, grooved axes, gouges, and polished slate articles were unknown to Lamoka. The burials associated with the Lamoka complex were of the flexed type (Ritchie, 1932). Grave goods were absent or scanty. Physically the Lamoka people belonged to a dolichocephalic group. Lamoka occupation was heaviest in western, central and southern New York. Lamoka-type points have been found in the lower levels of Hudson Valley sites and on certain sites farther east in lower New England (Ritchie, 1955). The main components of Lamoka are in western New York at Lamoka Lake, Geneva, and Scottsville. Ritchie’s (1944, 1955) second major division of the Archaic was the Laurentian which he believed to be later than Lamoka in the Northeast. In contrast to Lamoka he characterized Laurentian as a culture in which chipped-stone industries predominated over bone work. Since there was a wide range of projectile point types in Laurentian, Ritchie (1955, p. 6) believed that the fusion of discrete “local strains” probably occurred prior to the appearance of several focal complexes encompassed within the Laurentian tradition in the Northeast. Among the chipped-stone tools of Laurentian were several types of scrapers, bi-facially chipped knives, retouched flakes, and points which were prevailingly broad bladed, thick, with straight stem, or side or corner notches (Ritchie, 1955). The first trianguloid points supposedly made their appear- ance during Laurentian. A number of polished-stone tools have been attributed to Laurentian (Ritchie, 1955, p. 6). Included were celts, grooved axes, plano-convex adzes, beveled adzes, and gouges. The gouge came into use after the departure of the beveled adz and was considered as late in Laurentian (Ritchie, 1955, p. 6). Also included in the Laurentian inventory are plummets and atlatl weights in a variety of forms including notched, grooved, rectangular, and per- forated styles. Ritchie (1955, p. 6) believed that the bannerstone and the grooved ax diffused northward from the southeastern Archaic. Laurentian burials were often made in village refuse dumps but extended inhumations and cremations occurred along with flexed burials (Ritchie 1955, p. 7). Some grave goods, including red ocher, were often placed with the burial. Intentional burials of two breeds of dogs also occurred in Laurentian. The contents, distribution, and relationships of Laurentian to other groups in New York are complex. Ritchie’s (1944, 1955) attempts to draw together the traits of Laurentian tossed perhaps too many disparate things into one category. The excavated sites which produced Laurentian materials 218 Annals of Carnegie Museum vol. 35 also contained materials from later occupations. It was apparently not possible clearly to separate these different manifestations. Although most of the later materials were confined to upper levels, some late items were found mixed with Laurentian objects in lower levels. Most of the detailed information on Laurentian comes from three major sites— Robinson, Ober- lander (Ritchie, 1940) and Frontenac Island (Ritchie, 1944). All of these sites showed a mixture of materials. Several of the artifacts, such as winged bannerstones and gouges, that were assigned to Laurentian are typologically late in the Archaic. The presence of such objects at the excavated sites has done much to tag Laurentian as a late Archaic complex. This is a contention that must be more closely examined. MacNeish (1952, p. 47) stated that stratigraphic data were such that the Lamoka, Frontenac, and Brewerton foci could be arranged in chronological order and their genetic relationships could be determined. MacNeish (1952, p. 47) stated: “At the Lamoka Lake Site, Brewerton Focus artifacts appear near the top of the midden and above most of the Lamoka Focus materials (as well as under Vinette 1 pottery). Since the Frontenac Culture has traits of both Brewerton and Lamoka Lake, it is thought of by Ritchie as being chronologically between the Brewerton Focus and the Lamoka Focus.” Ritchie (1955, p. 3) stated that the priority of Lamoka over Laurentian had been stratigraphically established. In a search of the original site re- ports and subsequent publications, I have been unable to find demonstrable proof for such stratigraphy. Ritchie’s recent field work, yet unpublished, may hold the solution to this problem. Concerning the presence of Brewer- ton artifacts over Lamoka materials at Lamoka Lake just quoted above from MacNeish, Ritchie (1944, p. 307) was uncertain of the upper materials and stated that “the upper level complex from Lamoka Lake is, on the whole, still an enigma.’’ It is unlikely that Ritchie would have made such a statement if the identification of the upper level complex was positive for Brewerton. Ritchie (1944, p. 303) actually suggested relationships of the upper occupation materials with the “Early” Focus of the Coastal Aspect, but he also drew attention to the fact that the physical type differed. The Frontenac Focus was established on typological grounds as chrono- logically between Lamoka and Brewerton (Ritchie, 1944, p. 292. MacNeish, 1952, p. 47). No positive stratigraphy for the Lamoka and Brewerton ma- terials could be determined. Lamoka traits found at Frontenac included narrow-bladed stemmed and side-notched points, beveled adzes, choppers, antler pendants, notched pendant-like objects, and other more general traits. Brewerton traits at Frontenac included broad-bladed stemmed and side-notched points, plummets, bannerstones, perforated single-barbed har- poons, gouges, copper awls, and ground slate objects. Frontenac traits not common to Lamoka or Brewerton included elongated pendants of shell, receptacles of shell, new types of bone whistles, bird effigy combs, worked wolf mandibles, perforated elk and bear teeth, and new varieties of barbed harpoons. These last items are typologically late and bear some resemblance to objects in later Point Peninsula and Iroquois of New York. MacNeish (1952, p. 47) remarked that the presence of these items did not strengthen 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 219 the hypothesis of an early Brewerton group which moved into western New York, amalgamating with and replacing Lamoka. The sophisticated elements of Frontenac are more indicative of a late Archaic manifesta- tion than an early one. Therefore, it is difficult to see the Frontenac site as representing the contact of early Brewerton with the supposedly older Lamoka. If, however, Lamoka does not possess the antiquity and priority over Laurentian that has been attributed to it, the Frontenac situation of typological complexity and lateness becomes more understandable. The age of Lamoka becomes a crucial factor in northeast Archaic development and chronology. Since positive stratigraphic evidence for the priority of Lamoka over Laurentian has not been demonstrated and the typological picture is in- conclusive, it is necessary to turn to other types of evidence. Until many more Lamoka and Laurentian sites are investigated and reported in New York, this evidence must come from other areas. Frequent mention has been made of the resemblance of Lamoka to the Shell Mound culture of Kentucky (Ritchie, 1944, p. 322. MacNeish, 1952, p. 47). Unfortunately, these references have been vague and without sup- porting evidence, in May, 1956, with the co-operation of Dr. William S. Webb, I selected Lamoka and Laurentian-like points at random from In- dian Knoll and Annis Mound. The depth at which each of these points was found in the midden was then determined from the card catalogue and plotted on a graph. The results of this preliminary study reveal im- portant new information on the relationship of Lamoka and Laurentian to the Shell Mound culture. The Annis Mound is one of the oldest of the shell-middens. Materials from early to late Archaic have been found there. Laurentian-like small notched and expanded-stem points selected at random from this site were distributed in all levels from top to bottom. Of the 147 points in the sample, 43.53% were found above four feet, and 56.46% were found in the levels between four and eight feet. They were most abundant in the 6- and 6.5-foot levels. This distribution presents a picture of long and continued use of Laurentian-like points. Since such points were present at the low- est level at Annis Mound, considerable antiquity is established for the Laurentian-like points. The distribution of Lamoka-like points at Annis Mound contrasts sharply with the Laurentian distribution. Except for burial inclusions, where the digging for the burial had destroyed the natural stratigraphy, the Lamoka points are virtually all in the upper fourth of the midden. Of the 37 Lamoka points in the sample, 72.97% were found above the 1.5-foot level and 83.78% were above the 3.5-foot level and there were no Lamoka points in the lowest levels of the Annis Mound. At Indian Knoll the entire random sample of 21 Lamoka points was found above the four-foot level. Not one Lamoka point was found below four feet to the bottom of the midden at seven feet. Of the points present in the upper levels, 47.61% was found above the one-foot level. The distribution of Lamoka points at both Indian Knoll and Annis 220 Annals of Carnegie Museum vol. 35 Mound is most significant. Their placement in the upper levels at both of these can not be an accident. Careful field notes and records of depth for these points were made by Webb and his associates, and the methods and techniques of excavation were of the finest. The conclusion that Lamoka points are late in the Shell Mound culture of Kentucky seems in- escapable. MacNeish’s (1952, p. 47) statement that Lamoka has relationships with “the lower levels of the Annis Mound” does not agree with the facts. The relationship of Laurentian to Shell Mound culture has been ignored. Yet it is the Laurentian materials that are most abundant and widely dis- tributed within the Shell Mound culture. The presence of Laurentian points at the lowest levels at Annis Mound indicates that a Laurentian-like tradition was a basic element in Shell Mound culture. No such evidence has been presented for Lamoka. The radiocarbon dates for the Archaic in New York and Kentucky are of little aid in clarifying the chronology. Three dates have been obtained for Lamoka (Ritchie, 1951). Two of these dates were taken from charcoal samples at Lamoka Lake, while the third sample was removed from a hearth at Frontenac presumed to be of Lamoka origin. Sample 288 from Lamoka gave a date of 4369±200 years. As this sample was believed to be contaminated by recent carbon, another sample from a different hearth was dated at 5383±250 years. Ritchie (1951) accepted this later date as more nearly accurate for the Lamoka site. The sample from Frontenac Island which was presumed to pertain to the initial contact of Lamoka with Laurentian was dated at 4930±260 years. A date for Point Peninsula, which followed the Archaic in New York, has been given as 4400±260 years ago (Libby, 1952). If the Point Peninsula date is correct, the dated Archaic manifestations appear to be late as no great interval of time separates the two periods. Two recent dates from the Smith Farm site near Ellsworth Falls, Me., aid in the establishment of late Archaic and early ceramic contacts (Byers and Hadlock, 1955). One sample taken from middle levels considered as of Archaic origin was dated 3959±310 years. The other sample, taken from the uppermost level with Archaic materials, was dated 3350±4Q0 years. Immediately above the Archaic materials was a deposit comprising floors in which sherds of Vinette I pottery and later Point Peninsula pottery types were found. These dates would place the contact of the Archaic with the Point Peninsula culture 1000 years or more later in Maine than New York. The dates for the Shell Mound sites are even more perplexing than those for New York. The first dates for the Annis Mound were 4900 ±250, 5149±300 and 7374±500 years ago. (Webb, 1951). The earliest date, how- ever, was from a level above the two later dates. The shell sample dated from the three-foot level showed an age 2000 years greater than the shell sample from the six-foot level. Since the two samples were carefully collected from the same profile, Webb (1951, p. 30) believed that possibly the two samples had become interchanged in the laboratory. The date of 4900±250 was obtained from deer antler found at the 6.5-foot level. This date agreed favorably with the one of 5149±300 years for the same level. Two additional 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 221 dates were obtained later from the Annis Mound (Libby, 1953). Both of these dates were from composite samples of deer antler. The first was dated 4289 ±300 and was from levels 1.5, 2, and 2.5. The second sample was dated 4 3 33 ±4 50 and was from levels 5.5, 6, 6.5, and 7. While the first group of samples from Annis Mound was similar to those for New York Archaic, the second group was several hundred years later. The dates from the Indian Knoll site were also inconsistent (Webb, 1951. Libby, 1953). The first sample of antler tested from the one-foot level was dated 53O2±30O years (Webb, 1951). Later, another antler sample from the same one-foot level was dated 428 2 ±250 years (Libby, 1953). The inconsistencies of the dates from the Kentucky Shell mounds make them of little value and cast serious doubts upon the reliability of shell and antler for dating purposes. Differences of 1000 years for dates from the same level of a site make chronological comparisons of that site with more distant sites extremely difficult. Until many more dates can be ob- tained from sites and various levels within specific sites over wide areas of the Northeast, carbon dates can be only suggestive and not conclusive evi- dence for the age of a manifestation. Actual stratigraphy and comparison of materials associated with specific levels still remain the best methods of correlating chronology and relationships. In the Upper Ohio Valley there are no well defined manifestations which can be attributed to Lamoka. Stemmed points, plain adzes and beveled adzes similar to Lamoka have been found on many sites, but such items are usually associated with a predominantly Laurentian complex. There was no indication of the Lamoka complex at either the Dixon or Rohr sites. All materials from the lower Archaic levels at both of these sites are similar to those attributed to Laurentian. The small notched points, expanded- stem points, and bar atlatl weights found at Rohr are very similar to items found in the lower levels of the Annis Mound in Kentucky. The closest relationships of Lamoka to the Upper Ohio Valley are found in the “Panhandle Archaic” complex. Present in both of these complexes are numerous similar bone tools and plain and beveled adzes. The presence of beveled adzes in the “Panhandle Archaic” is the most distinctive link with Lamoka. Both Lamoka and “Panhandle Archaic” have individual traits which differ. Both, however, have relationships and contacts with Laurentian which is widespread over the Northeast. I believe that Lamoka and “Pan- handle Archaic” are distinctive regional complexes which developed upon a Laurentian base, subjected to differing influences from surrounding areas. Some of these influences were probably carried by new peoples. As shown by the presence of Lamoka-like points in the upper levels of the Kentucky Shell mounds, the typological similarities of “Panhandle Archaic” items with early ceramic manifestations, and the bone and polished stone traits shared by Lamoka and “Panhandle Archaic,” both complexes appeared late in the Archaic of the Upper Ohio Valley. Since several objections have been stated against Lamoka and the “Pan- handle Archaic” as representative of early Archaic in the Upper Ohio Valley, the origin, distribution, and relationships of Laurentian become of great importance. 222 Annals of Carnegie Museum vol. 35 Ritchie (1955, p. 8) did not believe that the Lamoka complex had de- veloped in situ from a Paleo-lndian base. The known Paleo-Indian assem- blages in the Northeast, from the Shoop (Witthoft, 1952), Reagen (Ritchie, 1953), and Bull Brook (Byers, 1954) sites represent a chipped-stone industry with fluted points, end- and side-scrapers based on trimmed flakes, gravers, and some knife forms usually from retouched flakes. None of these items has a counterpart in Lamoka. Conversely, Ritchie (1955, p. 8) also stated that “we seek in vain for a single Lamoka trait in the Paleo-Indian inventory.” The Laurentian, however, does possess traits that are similar to Paleo- Indian. Ritchie (1955, p. 8) stated that “in the Brewerton Focus of the Laurentian, on the other hand, a few elements of possible Paleo-Indian derivation occur, viz., uniface flake, end and side scrapers and knives, and a few lanceolate blades with parallel flaking.” All of these items, plus gravers, have been found on sites producing Laurentian materials in the Upper Ohio Valley. Gravers, end-scrapers, and side-scrapers were associated with the Laurentian points found in the lower levels of the Dixon and Rohr shelters. Two fluted points were found at the base of the Annis Mound in Ken- tucky (Webb, 1950). Although these points were believed to have been left on the site before the dwellers of the shell-midden arrived, there are traits within the midden which may have been derived from Paleo-Indian. Such traits are thumbnail end-scrapers, gravers, and end-scrapers. As noted pre- viously in this report, it is also within the lower levels of the Annis Mound that Laurentian-like points were most numerous. The Parrish site in Hopkins County, Ky., possessed an interesting mixture of Paleo-Indian and Archaic materials (Webb, 1951). During the excavation of this village a number of fluted points as well as typical Archaic artifacts were found. Items of particular interest are crude flake side-scrapers, thumb- nail end-scrapers, gravers, utilized flakes, and choppers. Webb (1951, p. 433) assigned these objects along with the fluted points to Paleo-Indian; but most of these items are also common in the Laurentian and occur on many of the Upper Ohio Valley sites assigned to that complex. The relationship of Laurentian to Paleo-Indian is far from being clearly understood. It seems unlikely that Laurentian was a direct development from Paleo-Indian. That the two cultures were in contact near the close of Paleo-Indian times is indicated by the traits they share and certain sites which possess remains of both. I believe the shared traits of Paleo-Indian and Laurentian give added support to the priority of Laurentian over Lamoka and other local complexes in the Northeast to which considerable age has been attributed. The origin of Laurentian has been sought by some workers in the sub- arctic zone of North America. This position was expressed by Spaulding (1955, p. 17) who stated: “From our standpoint, the important feature of the Laurentian is that it cannot be derived satisfactorily from the preceding cultures. It represents a new complex of elements associated with a new physical type in the northeastern part of our area. The gouges, ground slate points, and other elements point to a northern origin, and an actual movement of people from the Canadian forest seems to 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 223 be the most likely explanation. The ultimate origin of the Laurentian complex need not concern us here, but it is surely in some sense a representative of the north European-Siberian-North American cultural tradition of the boreal forest zone, and its appearance in the northeastern fringe of our area is a reflection of the proximity of that fringe to the northern forests. Indeed, certain parts of the area of Laurentian occupancy such as the Upper Peninsula of Michigan can be considered essentially of boreal forest type from the standpoint of ecology.” Ritchie (1955, p. 8) pointed out that it is not easy to find sites over the immense span of northern forests of Canada between the northeastern United States and the Alaskan bridgehead. Also, the recent views of Quimby (1953, p. 235-326) concerning the presence of ice of the Wisconsin glaciation in central and eastern Canada between 4500 B.C. and 2500 B.C. appear to rule out that area as a convenient passage for the Laurentian invasion of the Northeast. Since the Laurentian was already established in New York during that time, Ritchie (1955, p. 8) suggested that “we should seek a west to east route through the deciduous forest belt bordering the Great Lakes in Altithermal times.” I am in complete agreement with Ritchie’s suggestion that a west to east passage below the Great Lakes should be sought for Laurentian. Present evidence on the distribution of Laurentian material is indicative of such a passage. That this route was used by Laurentian is most clearly seen when we investigate the contents and relationships of local manifestations. Ritchie (1951, 1955, p. 6) remarked on the peculiar distribution within Laurentian of a dual category of ground slate tools comprising stemmed spear or lance heads, knives, and the ulo or semilunar knife. These forms occur separately or in various combinations on certain sites. Such items are known particularly from the Brewerton Focus, and the Vosburg Focus in New York, and the Vergennes Focus in Vermont (Ritchie, 1951. Bailey, 1939). The presence of slate items in Laurentian was considered to be by Martin, Quimby, and Collier (1947, p. 505) indicative of influences from the Dorset culture. Ritchie (1951, p. 50) suggested that the early occupation of the coniferous forest zone of North America by nomadic hunting bands from a similar Old World milieu was responsible for the establishment of these traits, probably first in Laurentian and later in Eskimo culture. This “proto-Laurentian” cultural stratum then underwent various regional elabo- rations. In Maine, for example, there was a rich efflorescence of ground slate points, gouges, adzes, and plummets which reach their apogee in the Red Paint Culture (Ritchie, 1951). In the Upper Ohio Valley slate objects are rare or absent on sites attrib- utable to Laurentian. When such items are found, they are usually present on sites adjacent to New York. Since slate objects are not common to the Laurentian sites of the Ohio Valley, it seems reasonable to consider such items as a specialization in the Northeast which came about through contacts with early Eskimo cultures. If such items were part of the early Laurentian when it passed through the Ohio Valley below the Great Lakes, it becomes difficult to explain their absence on the sites in that region. Willey and Phillips (1955, p. 751) summarized the Eskimo-Archaic problem as follows: “We have already pointed out the difficulty of applying our classificatory 224 Annals of Carnegie Museum vol. 35 criteria to cultural developments in the Arctic. Here we are dealing with cultural traditions so unique in total configuration as to make comparisons appear a trifle ridiculous. Nevertheless, with the possible exception of the Denbigh complex, as already noted, it is surprising how many of the criteria of our definition of Archaic are applicable to Eskimo cultures of the far north. This is particularly noticeable in the Dorset tradition of the eastern Arctic (Jenness 1930; Rowley 1940; Leechman 1943; Harp n.d.), whose relationship with northeastern Archaic is a perpetual subject of discussion (Hoffman 1952). But even such climatic Eskimo phases as Old Bering Sea (Collins 1937) and Ipiutak (Larsen and Rainey 1948) would, if subjected to the same unfavorable conditions for preservation, appear to be on the same general level of development as the well-known Archaic cultures further south. There is not only a basic conformity in stone and bone technologies but a great many shared elements of specific nature. Until fairly recently the tendency was to attribute the presence of such traits in Archaic contexts to ‘Eskimoan’ influence. Now, with the far greater time depth allowed to North American Archaic cultures, the problem of interchange between them and con- temporaneous Eskimo cultures has been immensely complicated, and it would be a rash student who would attempt to disentangle them.” The presence of copper tools in some New York Laurentian sites indi- cates important relationships with the Old Copper culture of Wisconsin. Typologically the copper gouge, adz, celt, and awl of Laurentian are similar to those of Old Copper (Ritchie, 1955). Until recently the Old Copper cul- ture was little more than a technological tradition. Work at the Osceola site (Ritzenthaler, 1946) and at the Oconto site (Ritzenthaler and Wittry, 1952) in Wisconsin has produced excavated specimens, and Wittry (1951) con- ducted a typological and distributional study of Old Copper artifacts in collections in Wisconsin. Popham and Emerson (1954) did a similar distri- butional study for Ontario and described the discovery of a cache of copper tools at Farquar Lake, Ontario. Two summaries of the Old Copper culture by Miles (1951) and by Wittry and Ritzenthaler (1956) appeared recently in American Antiquity. Wittry and Ritzenthaler (1956) noted the resemblance of many Old Copper traits to those of the Kentucky shell-middens and to those of the Brewerton and Frontenac foci of New York. The one bone whistle found with a burial at the Oconto site has a parallel in the Annis Mound in Kentucky. The burial complex of the Oconto site, however, has closer parallels with the New York Laurentian manifestations where cremations also are present, but there are differences in physical type between these early populations. Cremations are lacking in Lamoka (Ritchie, 1944, p. 394) and also in the Kentucky shell-middens, but they do occur in the lower levels of the Lauder- dale Focus in Alabama (Webb and Dejarnette, 1942). The radiocarbon dates of 5600±600 years and 7510±600 years (Libby, 1954, p. 740) for the Oconto site makes the Old Copper culture one of the earliest dated Archaic manifestations in North America and also indicates the earliest use of metals in the Americas and one of the earliest anywhere in the world. The large number and diversity of forms of the copper im- plements indicate that the Old Copper complex must have existed for a long time. Some Old Copper sites such as Osceola (Wittry and Ritzenthaler, 1956) possess lithic traits that show contact with Woodland groups. 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 225 Close resemblances of Laurentian and Old Copper are to be found not only in the use of copper tools and cremations but also in the use of certain lithic items. Although the chipped-stone complex of Old Copper is not well known, the few points from the early Oconto site are very similar in size and form to those of Laurentian. It seems entirely possible that the underlying base of Old Copper and Laurentian was the same. In Wisconsin, a specialized development of the use of copper tools began very early as a regional expression of this early Archaic base. At a later time the use of copper tools spread eastward around the Great Lakes into Ontario and New York and became a part of Laurentian. The relationships of the early lithic materials found in Canada to the Archaic of the Northeast are not well known. Lee (1955a, p. 4) stated that Lamoka traits for early Archaic are “not at all applicable to Ontario.” At the Sheguiandah site on Manitoulin Island, Lee (1955b, p. 71) noted the presence of materials which he considered to be an early variant of Lauren- tian. Ridley (1954, p. 43) made the following correlation for materials at the Frank Bay site, Ontario: “Partial correlation of the Frank Bay Mattawan complex is geographically wide- spread but exact duplication is unknown. Archaic material similar to that from western New York is found elsewhere on the perimeter of Lake Nipissing, is re- ported at several stations many miles north of Lake Nipissing, and is well repre- sented by sites and surface collections over much of southern Ontario. Nevertheless, though Mattawan and eastern Archaic share many features, differences are also great, and the two are clearly not the same. Neither does the Mattawan complex appear to be proto-Archaic, nor a regional variation of eastern Archaic, but seems to be considerably older than this general phase. Similarly, there are suggestive parallels with Paleo-Indian sites; for example Reagan (Ritchie 1953), Shoop (Witthoft 1952) and Brohm (MacNeish 1952) with materials from Quebec (Rogers and Rogers 1950), and with arctic stations, for example, Dorset (de Laguna 1946; Wintemberg 1939). Once again differences are obvious, and in any event too little is known of the Mattawan complex to justify interpretation at this time. Recently Paul Sweetman and I discovered a site 20 feet above the shore of Lake Temis- kaming, well to the north of Lake Nipissing. In a dark top stratum, artifacts tentatively identified as Archaic, resembling Atlantic states Archaic, and Indian Knoll, overlay a deep deposit of yellow sand and alluvial gravel. In the sand, and well below the gravel, were small hearths and 8 Mattawan type artifacts.” In the MacKenzie Valley (MacNeish, 1954, p. 234-253), stemmed and notched projectile points comparable with Eastern Archaic forms have been found in association with specialized tools such as burins, microblades, and polyhedral cores. Wettlaufer (1955, p. 91-113) illustrated many points from the Mortlach site in Saskatchewan which resembled those typical of Laurentian. There can be no doubt that typologically similar points occur both at the Laurentian sites of the Northeast and Ohio Valley and at several sites in Canada. Until many early Archaic sites are known in the Northeast and Ohio Valley, it will be difficult to demonstrate what relationships may exist be- tween these sites and those in Canada. One of the most interesting Archaic sites in the Ohio Valley is the McCain site in Dubois County, Ind. The work at this site was conducted by Rex K. 226 Annals of Carnegie Museum vol. 35 Miller (1941) from 1929 through 1934. This work and the excellent report which followed were completed before the Archaic manifestations in Ken- tucky and the Northeast were well known. As a result, important relation- ships of the McCain site materials with Laurentian have been overlooked. Through the courtesy of Glenn A. Black of the Indiana Historical Society, all the lithic specimens from the McCain site were made available to me for study and revaluation. The information which follows is based on my recent study of the materials as well as Miller’s (1941) report. The McCain site is located on an ancient terrace along the East Fork of White River in Boon Township, Dubois County, Indiana. This terrace has been heavily cut by erosion and the river has shifted its channel until now it is some distance away from the site. Miller’s (1941, p. 10-16) excavations revealed that the site was covered in many places by an occupation debris of humus-laden soil containing ash, charcoal, bone, shell, burned stones, and artifacts to a depth varying from 2lA to 314 feet. Along the slopes of the terrace were refuse dumps as deep as 414 feet. These dumps contained great quantities of mussel shells. No evi- dence of house sites, permanent fireplaces, storage or refuse pits were found. Large quantities of ash and burned stones indicating temporary fires were found scattered throughout the debris. Artifacts were found in both the village debris and the dump areas. Twenty flexed burials were found in shallow pits in the debris. Many fine bone objects were found at the McCain site. Their state of preservation, due to the large amounts of shell in the debris, was excellent. Included among the bone items are antler punches and drifts, awls of deer and bird bone, bone scrapers and knives, and fish-hooks. However, the most interesting bone objects are a number of bone pins with incised geometric decorations. Some also have notches and many have a perforation at one end. These and the other bone objects were illustrated by Miller (1941, p. 35-47). The decorated bone pins are similar to those from Lamoka (Ritchie, 1932, p. 106), and they are especially similar to the incised bone items from the Annis Mound in Kentucky (Webb, 1950, p. 325). Among the polished-stone tools of the McCain site, the 31 full-grooved axes were the most important and the only large cutting tool except for four crude, pecked-stone choppers. Three of the full-grooved axes were made of limonite. Other items of stone included 56 hammerstones and 13 mullers. No atlatl weights were found. Most of the polished-stone tools, particularly the full-grooved axes, are similar to those of the Kentucky Archaic and other scattered Archaic sites throughout the Ohio Valley. The chipped-stone tools of the McCain site present many interesting similarities to tools from other Archaic sites in the Ohio Valley. Since Miller (1941) touched only lightly on these items in his report, they are described more fully here. All the chipped-stone tools were separated into types. Some items were difficult to classify because of their indefinite form, and could logically be placed in more than one category. This material may be summarized as follows: Small side-notched points. The 85 small side-notched points are variable 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 227 in outline but all possess crudely chipped side notches. Many of these points have concave bases, and some have projecting ears. Basal thinning occurs on several specimens. The flaking is crude and the material is local cherts. They range in length from % inch to 2 inches. These small points are similar to those found on the Laurentian sites of the Upper Ohio Valley and Northeast. They are especially like those found at the Dixon and Rohr shelters and at the Shriver site (Stewart and Dragoo, 1954). A close resem- blance can be seen in these points to those from the Annis Mound in Ken- tucky (Webb, 1950) and from the Faulkner site in Illinois (Cole, 1951, p. 216). (See Fig. 29.) Stemmed points. There are 51 points with stems. Most of these points have a weak shoulder and short, parallel-sided, rounded base. They range in length from \x/% inches to 3]4 inches. The larger of these points are long and slender, while the smaller ones tend to be more broad in relation to length. Chipping is generally crude except for a few specimens which were finely executed. Many of these points have bases which are thinned. The material is variable with local coarse-grained cherts predominating. The best made and largest of these points are of Harrison County, Ind., flint. In form many of these points are similar to those in the Archaic levels of the Dixon and Rohr sequence. Not only do they resemble the Lamoka points of New York (Ritchie, 1944, p. 299), but also similar points are found on Laurentian sites in the Upper Ohio Valley and Northeast. (See Fig. 30.) Expanded-stem points. There are 39 points with stems formed by cutting crude side or corner notches into the point. These notches, however, are not well defined as in the points assigned to the side-notched and corner- notched categories. These points resemble the stemmed points both in the crude chipping and the materials used. Some show basal thinning. They range in length from l1/^ inches to 3 inches. These points are very similar to those found on the Upper Ohio Valley Laurentian sites. (See Fig. 31.) Large side-notched points. There are 19 side-notched points ranging in length from 2 inches to 23A inches. These points are similar to the small side-notched points except that they tend to be more slender, and finer flint, such as that from Harrison County, Ind., was more commonly used. (See Fig. 32.) Corner-notched points. There are 22 corner-notched points ranging in length from one inch to three inches. The notches in these points are deeply cut, and the width of the point is wide relative to the length. The chipping was finely executed and flint from Harrison County, Ind., and Flint Ridge, Ohio, was most commonly used. Some local materials, however, were used for a few of the specimens. Although corner-notched points do occur in the Archaic, they are commonly associated with the later Woodland cultures. (See Fig. 33.) Miscellaneous points. This group was composed of a small number of points from the McCain site which do not fit well into the above categories. Among these points are two large side-notched blades with deeply cut notches and wide, slightly rounded bases which have been ground smooth. The cutting edges are beveled. There are also one large and two small side- 228 Annals of Carnegie Museum vol. 35 Fig. 29. McCain site. Small side-notched points Fig. 30. McCain site. Large stemmed points 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 229 Fig. 32. McCain site. Side-notched points and knives Fig. 31. McCain site. Expanded-stem points 1 230 Annals of Carnegie Museum vol. 35 notched points with deep notches and straight rectangular bases. There is also one large, wide, corner-notched point with deep notches and serrated edges. All of the above points are common to Middle Woodland sites. Their distinctiveness at McCain suggested to me that a later component may be present. (See Fig. 33.) The remaining miscellaneous items included 2 lozenge-shaped blades, 1 knife-like blade, 3 large blades with shallow notches, 2 small triangular points, and 18 broken points. Drills.. There are 19 complete drills and 10 fragments of drills in the McCain site collection. These drills were classified as follows. Ten expanded- base drills made of varied materials, two drills with side-notched bases, one straight drill, one expanded-center drill, and five crudely chipped drills with triangular cross-section and expanded bases. The 10 fragments con- sisted of tips and were unclassifiable as to type. All of these drills are similar to those found in Archaic manifestations both in the Northeast and Kentucky. (See Fig. 33.) Scrapers. The most distinctive scrapers in the McCain collection are the 71 end-scrapers. The scrapers of this group are either stemmed, side notched, or corner notched. Only one scraper is of the thumbnail type. Many of these scrapers appear to be reworked points. The material varies from crude local cherts to fine Harrison County, Ind., flint. These scrapers are similar to those from Indian Knoll in Kentucky (Webb, 1946, p. 263) and Faulker in Illinois (Cole, 1950, p. 216). (See Fig. 34.) There are several side-scrapers of an unspecialized form. Some of the crude blanks were also used as scrapers. Blanks. There are many roughly percussion-chipped blanks of various sizes in the McCain collection. Most of these show no secondary usage, but a few specimens have chipped and battered edges. The importance of the McCain site lies in its similarities to several Archaic manifestations from widely separated areas of the Ohio Valley. The many expanded-stem and side-notched points are particularly similar to those of the -Laurentian sites of the Northeast and the Upper Ohio Valley. The full-grooved axes, stemmed points and end-scrapers, however, are more closely related to the Kentucky Archaic and Faulkner of Illinois. The flexed burials and the quantities of shell in the village follow the pattern of the Kentucky Archaic. Again in the McCain site as in the Annis Mound there is evidence of an early Laurentian lithic tradition in the Ohio Valley. In Indiana there are many sites on which Archaic materials are found. The best known of these sites are the shell-middens along the Ohio River and the lower Wabash River (Adams, 1949. Dragoo, 1955. Kellar, 1956). Only surface collections are known for most of these sites. In general, they are similar to the Kentucky Archaic sites, but many of the point types from these sites are also similar to those of the Upper Ohio Valley and northeast Laurentian sites. The chronological placement of most of these sites within the Archaic is unknown. Black’s (1946) report on the Cato site in Pike County, Inch, indicated the interaction of a late Archaic mani- festation with Adena. The McKinley site near Noblesville, Inch, recently 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 231 Fig. 33. McCain site. Drills and corner-notched points Fig. 34. McCain site. End-scrapers 232 Annals of Carnegie Museum vol. 35 excavated by Downey Raibourn for the Indiana Historical Society, appears to be of an earlier date. Less well known than the large shell-middens, are many small Archaic sites scattered over the entire state of Indiana. These small sites usually do not contain shell debris, but their lithic inventory is similar to that of the shell-midden sites. Many of these sites are located on terraces: which are now at a considerable distance from the nearest streams. One of the pressing needs of midwestern archeology is the thorough investigation of these small Archaic sites which have often been ignored or mislabeled “Woodland Campsite” in the reports. The Archaic manifestations of Ohio are even less well known than those of Indiana. Scant attention is given to the Archaic since so many productive burial mounds are present to lure the archeologists and collectors. Morgan (1952, p. 84) pointed out the presence in Ohio of materials similar to those of Indian Knoll, and suggested that the area was occupied by related peoples during the Archaic. Laurentian-like materials are also present in many of the collections from northern Ohio (Smith 1954, p. 9-11). Surface collections recently made by Charles Sofsky of Warren, Ohio, and John Zakucia of Youngstown, Ohio, from sites in northeastern Ohio contain many Laurentian items. Investigations are now underway at many of these sites and reports of the findings should appear in the near future. The materials from the Faulkner site in southern Illinois bear many re- semblances not only to Laurentian but also to the Kentucky Archaic (Cole and others, 1951). Many of the notched and expanded-stem points, drills, and end-scrapers are similar to those of Laurentian sites in both the Upper Ohio Valley and New York. The notched and stemmed scrapers and the straight-stemmed points are similar to those of Indian Knoll (Webb, 1940) and the McCain site (Miller, 1941). The concave-based points of Faulkner are within the range of some Paleo- Indian fluted points, and their specific resemblance to points of the Par- rish site in Kentucky was noted by Cole (1951, p. 216). Again, as with both Laurentian and the Kentucky Archaic, there are elements in Faulkner which may have been derived from the earlier Paleo-Indian. The work conducted by the Illinois State Museum at the Modoc Rock Shelter in southern Illinois added much to our knowledge of the age of cer- tain tools associated with the Archaic (Fowler and Winters, 1956). The radiocarbon dates obtained from charcoal found at this site range from B.C. 2665±300 years to B.C. 7922±392 years for Archaic deposits. The major cultural manifestation at the Modoc Rock Shelter was Archaic and was found from a depth of 5 feet down to 21.5 feet. The upper five feet contained pottery that indicated middle to late Woodland occupation of the site. Within the Archaic zone there are changes through time that in- dicate the increasing complexity of the culture. Among the changes were: 1. a change from predominantly side-notched, slender projectile points to a variety of forms that are predominantly stemmed, notched, and broad bladed, and, 2. an increase in the number and variety of chopping and cut- ting tools and polished-stone artifacts. Fowler and Winters (1956, p. 44) 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 233 tentatively set apart the lowest zone of the site (21.5 to 26.5 feet) from the higher layers. They believed the one broad tapered-stemmed projectile point found in this lowest zone was similar to those found in the lowest portions of the Hidden Valley Shelter in Missouri (Adams, 1941). Also, these points are similar to the Gypsum Cave points of Nevada (Harrington, 1933) that were associated with extinct fauna and have been dated at a time contem- porary with the lowest zone of the Modoc Rock Shelter. The projectile points from the Archaic zone of the Modoc Rock Shelter possess many resemblances to points of the Kentucky Archaic and of Laurentian. The small side-notched points (Fowler and Winters, 1956, p. 18, 22) are especially similar to those of the Upper Ohio Valley Laurentian as found in the Dixon and Rohr sequence. The choppers, side-scrapers, and end-scrapers are also to be found in Laurentian manifestations in the Upper Ohio Valley. The large stemmed points and notched end-scrapers of Modoc are more at home in the Kentucky Archaic and at sites like McCain in Indiana (Miller, 1941). The radiocarbon dates for the Archaic at Modoc Rock Shelter push cul- tures of this period back to at least 6200 B.C. Such a date is indicative of the antiquity that can be assigned to other Archaic manifestations in the Ohio Valley. The finding of Laurentian-like items at Modoc is again evidence of the presence of an early basic Archaic culture below the Great Lakes upon which Laurentian developed in the Upper Ohio Valley and the Northeast. Outside of the Ohio Valley and the Northeast, the relationships of Laurentian with other Archaic manifestations of the eastern United States are not well known. In the Carolina Piedmont, Coe (1952, p. 306) considered the Badin culture as similar to the Indian Knoll culture of Kentucky. He considered the correlation of traits so great that a direct connection between the two groups was postulated. Since the Badin culture appeared in the Piedmont about the same time the Kentucky Archaic was being displaced by Adena, it is late in the Archaic sequence. Within the Badin culture are some items such as projectile points, drills, and scrapers (Coe, 1952, fig. 163) which are similar to Laurentian forms, but these similarities probably rep- resent the basic underlying relationships of Kentucky Archaic and Lauren- tian. The Early Archaic of the Piedmont, as manifested by the Guilford Focus (Coe 1952, p. 304) appears to be unrelated either to the Badin culture, Kentucky Archaic, or Laurentian. In the lower Susquehanna Valley and in the Potomac Valley, the Laurentian did not gain a major foothold. Some minor thrusts of Lauren- tian into these areas did occur. A few Laurentian materials are present in private collections from sites along the upper Potomac and the upper Susque- hanna. The early Archaic manifestations of these areas are not well known. Recent work by John Witthoft of the Pennsylvania State Museum indicated the presence of an early chopping-tool assemblage which may have consider- able antiquity. The late Archaic is best represented by the broad spearpoint tradition and the use of steatite pots (Witthoft, 1953, p. 4-31). Contact of the “Transitional Cultures” with Laurentian did occur on the eastern 234 Annals of Carnegie Museum vol. 35 margins of the Upper Ohio Valley. Witthoft (1953, p. 25) suggested that the beginnings of pottery in the East was perhaps derived from the soap- stone vessel tradition of the “Transitional Cultures.” Graham Cave (Logan, 1952), located in Logan County, Missouri, is a stratified site that was occupied primarily during the Archaic, but earlier Paleo-Indian materials were found in the lowest levels. A distinctive feature of the site is the presence of a series of lanceolate projectile points. There are two major varieties of these points. The first is characterized by short channel flakes on one or both faces, by grinding on the sides which pro- duced a slight expansion at the base, and by a concave base which has slightly flaring “ears.” The second variety has a flat base and is propor- tionately shorter and broader. While both varieties are more abundant in the lowest level, they are in later levels. Logan (1952, p. 67-68) interpreting this material as a fusion of cultural traditions, stated: “The Archaic horizon of Central Missouri is not the product of one stream of cultural influence. Traits from several different sources have fused to produce the Archaic culture seen in Graham Cave. The Archaic horizon in level 6 is a series of Eastern Archaic elements injected into a previously existing complex which has the appearance of having had a western origin. This original complex included the concave based points with flutes and basal grinding reminiscent of the Clovis Fluted points. Since a Pleistocene or near Pleistocene antiquity for this level seems improbable, there must have been an antecedent to Graham Cave elsewhere in the region.” Logan (1952, p. 68-69) also pointed out that “the Archaic complex of Graham Cave especially from levels 4 and 5 on is essentially a complex with an eastern orientation” and that “from level 5 up, Graham Cave is comparable to the southeastern material.” Implied in Logan’s remarks was the belief that a Paleo-Indian horizon was basic in this region and that an Archaic complex from the East interacted with and finally completely absorbed this Paleo-Indian tradition. Baerreis (1955, p. 7) presented an alternative hypothesis stating that “a Paleo-Indian tradition was involved in a pattern of acculturation with an already resident archaic tradition.” This hypothesis not only has merit for the Graham Cave situation, but also for other areas in the Eastern United States. Since the bulk of our information pertains to sites with already well developed Archaic complexes, one of the basic needs for the understanding of the origin and spread of the Archaic is the discovery and definition of early Archaic manifestations. The lower time limits and relationships of the Archaic with Paleo-Indian cultures remain one of the most interesting problems in American archeol- ogy. Direct contacts for Archaic and Paleo-Indian occurred in both the Kentucky Archaic and at Graham Cave. The presence of scrapers and other forms in Laurentian similar to those known for Paleo-Indian are important evidence of early contacts. The small bifurcated points found in the bottom levels of the Rohr Shelter and on many Archaic sites are sug- gestive of points associated with the Pinto Basin Culture in the West. Many of the crude choppers, scrapers, and gravers of the Archaic may be related to an early lithic complex different from Folsom Paleo-Indian, but with as great, if not greater, antiquity. Until extensive work over large areas is 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 235 concentrated on early sites, the beginnings of the Archaic will remain only speculative. SUMMARY AND CONCLUSIONS The Archaic cultures of the Upper Ohio Valley and their relationships to surrounding Archaic manifestations in the Eastern United States are summarized as follows: 1. In the lowest levels of the Dixon and Rohr shelters was found an early Archaic culture which had as its major tools small expanded-stem and notched points, small end-scrapers, side-scrapers, utilized flakes, small bar atlatl weights, crude hammerstones, and pitted stones. Many small open sites, usually located on prominent hilltops near streams or springs such as the Shriver site and the Wheeling College site, had a similar chipped-stone complex plus a more extensive inventory of polished-stone tools such as full-grooved axes and adzes. The economy of the people living on these sites was based on hunting, fishing, and gathering. Deer, elk, and certain small game were major items on the menu. The presence of pitted and abraded stones on many of these sites indicated that wild vegetal foods were also an important part of the diet. In contrast to the lower Ohio Valley and Tennessee Valley Archaic sites where shellfish were of great importance as a source of food, the Upper Ohio Valley Archaic sites only rarely contain shellfish remains. The rock shelter sites were probably used most extensively during the winter months while the open sites were used during months of good weather. No evidence of substantial dwellings of any type has been found on any of the Upper Ohio Valley Archaic sites. 2. The material culture of these early Archaic sites has been compared most often with the Lauren tian of the Northeast. The materials are similar but important differences do exist. The early Upper Ohio Valley Archaic sites contain items, such as grooved axes and bar atlatl weights, which appear to be earlier than the adzes and winged atlatl weights associated with New York Laurentian. The materials from the lowest levels of the Dixon and Rohr sequence are representative of an early Archaic tradition which de- veloped upon an early base similar to, and probably ancestral to, the Laurentian of the Northeast. Tentative observation of the materials from Archaic sites in the Upper Ohio Valley has indicated that the materials tend to be progressively later from south to north with the more northerly Ohio Valley sites most closely related to the Laurentian of New York. Since the term Laurentian has been in the literature for several years, its use has been extended in this article to cover related Archaic manifesta- tions, some of which are much earlier than the New York Laurentian, in areas surrounding the Northeast. Some of my colleagues may object to this extension of Laurentian to cover these new manifestations; however, I be- lieve at the present time the use of Laurentian or Laurentian-like to be preferable to the addition of a new term to the literature. 3. The Upper Ohio Valley Laurentian manifestations are related to an early Archaic culture which spread over much of the Ohio Valley and into the Northeast. Many Laurentian elements are found in the early develop- 236 Annals of Carnegie Museum vol. 35 ment of the Kentucky Archaic. At the Annis Mound and Indian Knoll of the Kentucky Archaic, Laurentian-like points were found from the lowest to the highest levels. Important basic elements in Laurentian are also present in such sites as McCain in Indiana, Faulkner in Illinois, and in the Old Copper complex of Wisconsin. Laurentian-like materials are present in many surface collections from Archaic sites scattered throughout the Ohio Valley, southern and northwestern Canada, and portions of the Mississippi Valley. 4. The early basic Laurentian culture entered the Ohio Valley from northwest of the Great Lakes, spread throughout the area below the Great Lakes and into the Northeast while glaciers still covered much of central Canada. With the glacial retreat, Laurentian spread northward into central and eastern Canada. Some ancestral Laurentian groups were contemporane- ous with late Paleo-Indian peoples and contacts between these two cultures undoubtedly occurred. (See Fig. 35.) 5. Having established itself in the Ohio Valley and the Northeast, the basic Laurentian culture underwent many local modifications. In the North- east, plain and beveled adzes and slate tools became part of the complex. The use of copper for tools was developed early by peoples possessing a Laurentian-like culture in Wisconsin. In Kentucky and surrounding areas, Laurentian elements were important early components in the formation of the extensive shell-mound Archaic. Time, isolation, dissimilar natural en- vironments, different contacts with other groups, and varied responses to their problems by the people in the widely separated areas were important factors in the development of distinctive complexes in late Archaic times. 6. The Upper Ohio Valley Laurentian sites during late Archaic times had varied contents depending upon their nearness to other major Archaic complexes in surrounding areas. The northern sites have traits, such as plain adzes, gouges, and winged atlatl weights, which are common in the New York Laurentian. The southern sites have traits, such as the full- grooved ax and bar atlatl weights, which link these sites with the Kentucky Archaic. The sites on the northern and eastern fringes of the Upper Ohio Valley near the Susquehanna and Potomac drainages were greatly influenced by the “Transitional Culture”. Found on these sites are large tapered-stem blades, semi-lozenge points, and steatite vessel fragments. Studies now in progress indicate that the “Transitional Culture” was more important on the late Archaic scene in the Upper Ohio Valley than was previously sus- pected. 7. Beginning perhaps during middle Archaic and extending into late Archaic times in the Upper Ohio Valley proper, the “Panhandle Archaic” developed. Among the most important tools of the “Panhandle Archaic” are the grooved adz, stemmed point, lanceolate blade, and crescent banner- stone. Outside of this area, the grooved adz is best known from late Archaic sites in New England and along the east coast. What connections, if any, existed between these widely separated areas remain unknown. The bone items of the “Panhandle Archaic” are similar to those of Lamoka of New York and also to Kentucky Archaic. The sites of this complex are the only 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 237 Fig. 35. Probable route of Laurentian culture into the eastern United States 238 Annals of Carnegie Museum vol. 35 sites with shell-midden deposits in the Upper Ohio Valley. The peoples responsible for the “Panhandle Archaic” were certainly in contact with Laurentian culture and may have participated in the development of Early Woodland cultures in the same area. 8. The Lamoka complex of New York presents questions difficult to answer at this time. Its existence early in west-central New York is un- deniable, but outside of that area the Lamoka as a total complex does not seem to exist. Individual traits such as the beveled adz and occasionally the typical narrow, stemmed points of Lamoka occur on sites in the Upper Ohio Valley, but these sites always possess predominantly late Laurentian traits. Some Lamoka-like points have been found in the Kentucky Archaic at Annis Mound and at Indian Knoll. At both these sites the Lamoka points were found only in the upper levels in contrast to the bottom-to-top distribution of Laurentian points. Only continued research will tell us whether the Lamoka is only a regional specialization of the basic Archaic culture which was also ancestral to Laurentian in that area, or whether Lamoka had en- tirely different origins. 9. The Adena peoples of the Early Woodland period seem to have pushed Archaic peoples with Laurentian culture into less accessible areas away from the Ohio River Valley proper. Within the Middle Woodland complexes of the Upper Ohio Valley are a number of Laurentian traits, espe- cially point types, which indicate that Laurentian influences persisted long after the introduction of pottery and different economic pursuits. CARBON-14 DATES FROM ROHR SHELTER While this report was in press the following radiocarbon dates for the Rohr Shelter were received from Dr. Edward S. Deevey, Yale University, Geochronometric Laboratory, New Haven, Connecticut (Personal communi- cation from Deevey to Dragoo, April 18, 1958): Charcoal from Rohr Shelter (39°35' N. Lat., 79°50/ W. Long.), Monon- galia County, West Virginia. The samples span the history of Archaic culture at the site, where it is represented by the lower two-thirds of the deposit, overlain by pottery bearing layers. The underlying yellow sand is intrpreted as fluviatile, laid down when Bee Run, now many feet below the shelter, was cutting the rock formation in which the shelter is found. Col- lected in 1955 and submitted by Don W. Dragoo and Albert Bauer. Y484. Rohr Shelter, Late Archaic 2200 ±60 From hearth in block A3, 18 inches below surface, sample 46 Mg Cl, in transition zone from Late Archaic to Early Woodland. Y485. Rohr Shelter, Middle Archaic 1940±70 From hearth in block A3, 24-26 inches below surface, sample 46 Mg 9-C2, about midway in Archaic deposit; the culture materials are similar to those of Laurentian culture in New York state. Y486. Rohr Shelter, Early Archaic 5310±90 From hearth in block A2, 32-36 inches below surface, sample 46 Mg 9-C3, just above sterile yellow sand; the artifacts are typologically early Archaic. 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 239 Comments: The difference between the ages of Y484 and Y485 is not sig- nificant. The three samples give a satisfactory dating of Archaic culture in the Upper Ohio Valley, from its inception through its transition to the pottery making Early Woodland culture. The difference between the ages of Y484 and Y485 may best be explained on the basis of contamination with recent carbon from fine rootlets which permeated all the samples from the shelter. As many of these rootlets as possible were removed from the samples but because of their smallness many still remained. The presence of more recent carbon in sample Y485 than in sample Y484 may account for the reversed order of these dates. The presence of modern carbon in all the samples would tend to make the dates more recent than perhaps they are in actuality. The date of 5310±90 for the low level at the Rohr Shelter is very close to the early dates obtained for the Archaic shell mounds along the Green River in Kentucky and for the early Lamoka culture in New York state as mentioned in the preceding section of this report. The 5310±90 date may be interpreted as a conservative date for the presence of proto-Laurentian culture in the Upper Ohio Valley and would indicate that Laurentian was present in this area earlier than its appearance in the Northeast. REFERENCES Adams, Robert McCormick 1941. Archaeological investigations in Jefferson County, Missouri, 1939-40. Transactions of the Academy of Science of St. Louis, v. 30, p. 151-221. Adams, William R. 1949. Archaeological notes on Posey County, Indiana. Indiana His- torical Bureau, 81 p. Indianapolis. Anonymous 1943. First archaeological conferences on the Woodland pattern. American Antiquity, v. 8, no. 4, p. 393-400, Menasha, Wis. Baerreis, David A. 1955. The Archaic as seen from the southwestern periphery. Mimeographed paper presented to Society for American Archaeol- ogy conference, Bloomington, Ind., 1955. Bailey, John H. 1939. A ground slate producing site near Vergennes, Vermont. Bulle- tin of Champlain Valley Archaeological Society, v. 1, no. 2, p. 1-29. Fort Ticonderoga, New York. Bennett, Wendell C. 1948. A reappraisal of Peruvian archaeology. Memoirs of the Society for American Archaeology, v. 8, no. 4, p. 1-7. Menasha, Wis. 1953. New world cultural history; South America. In “Anthropology today,” University of Chicago, p. 211-225. Black, Glenn A. 1946. The Cato site, Pike County, Indiana. Proceedings of the Indiana Academy of Science, v. 55, Indianapolis. 240 Annals of Carnegie Museum vol. 35 Butler, Mary 1936. Archaeological problems in Erie County, Pennsylvania. Penn- sylvania Archaeologist, v. 6, no. 2, p. 27-30. Milton, Pa. 1939. Three archaeological sites in Somerset County, Pennsylvania. Pennsylvania Historical Commission, Bulletin no. 753, p. 1-79. Harrisburg. Byers, Douglas 1954. Bull Brook— a fluted point site in Ipswich, Massachusetts. American Antiquity, v. 29, no. 4, p. 343-351. Salt Lake City. Carpenter, Edmond S. 1942. Archaeological reconnaissance in the Upper Ohio Valley. Penn- sylvania Archaeologist, v. 12, no. 1, p. 20-23. Milton, Pa. 1951. Tumuli in southwestern Pennsylvania. American Antiquity, v. 16, no. 4, p. 329-346. Menasha, Wis. Chamberlain, Thomas C., and Rollin D. Salisbury 1921. Introductory geology. United States Geological Survey. Wash- ington. (Water supply paper 234.) Coe, Joffre 1952. The cultural sequence of the Carolina Piedmont. In “Archeol- ogy of eastern United States,” p. 301-311. University of Chicago Press, Chicago. Cole, Fay-Cooper, and others 1951. Kincaid; a prehistoric Illinois metropolis. 385 p. University of Chicago' Press, Chicago. Collins, Henry B. 1937. Archaeology of St. Lawrence Island, Alaska. Smithsonian Mis- cellaneous Collections, v. 96, no. 1, Washington. Corliss, Frank R., Jr. 1951. The finding of an unusual cache at an Archaic site. Archeolog- ical Newsletter, no. 4, p. 1. Carnegie Museum, Pittsburgh. 1954. Archeology in Garrett County, Maryland. Archeological News- letter, no. 9, p. 6-12. Carnegie Museum, Pittsburgh. Cross, Dorothy 1941. Archeology of New Jersey. Archeological Society of New Jersey, v. 1, 271 p. Trenton. deLaguna, Fredrica 1946. The importance of the Eskimo in Northeastern archeology. In “Man in northeastern North America.” Papers of Robert S. Peabody Foundation for Archeology, v. 3, p. 106-142. Andover. Dragoo, Don W. 1953. Excavations at Watson site, Hancock County, West Virginia. Bulletin Eastern States Archeological Federation, no. 13, p. 9. 1955a. Excavations at Johnston site, Indiana County, Pennsylvania. Pennsylvania Archaeologist, Memoir no. 1, p. 1-56, Milton, Pa. 1955b. Archaeological survey of Gibson County, Indiana, p. 1-39. In- diana Historical Bureau, Indianapolis. 1956. Excavations at the Watson site, 46Hk34, Hancock County, West 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 241 Virginia. Pennsylvania Archaeologist, v. 16, no. 2, p. 59-88. Milton, Pa. Fairbanks, Charles 1942. The taxonomic position of Stallings Island, Georgia. American Antiquity, v. 7, no. 3, p. 223-231. Fetzer, Elmer, and William J. Mayer-Oakes 1951. Excavation of an Adena burial mound at the Half-Moon site. West Virginia Archeologist, no. 4, p. 1-25. Moundsville. Fowler, Melvin L., and Howard Winters 1956. Modoc Rock Shelter. Illinois State Museum Report of Investi- gations, no. 4, Springfield. Griffin, James B. 1942. Adena pottery. American Antiquity, v. 7, no. 4 p. 334-358. Menasha, Wis. 1943. Adena village site pottery from Fayette County, Kentucky. Uni- versity of Kentucky Reports in Anthropology and Archaeology, v. 5, no. 7, p. 666-672. Lexington. 1946. Cultural changes and continuity in eastern United States archeology. In “Man in northeastern North America.” Papers of Robert S. Peabody Foundation for Archaeology, v. 3, p. 37-95. Andover, Mass. 1952. Culture periods in Eastern United States archeology. In “Arche- ology of eastern United States,” p. 352-364. University of Chicago Press, Chicago. Hadlock, Wendell S., and Douglas Byers 1955. Radiocarbon dates from Ellsworth Falls, Maine. American An- tiquity, v. 21, no. 4, p. 419-420. Salt Lake City. Harp, Elmer, Jr. 1952. Cultural affinities of the Newfoundland Dorset Eskimo. Ph.D. Dissertation at Harvard University, Cambridge, Mass. Harrington, M. R. 1924. The Ozark Bluff-Dwellers. American Anthropologist, v. 26, no. 1, p. 1-21. Menasha, Wis. Hennen, Ray, and David B. Reger 1914. West Virginia Geological Survey— Preston County. Wheeling News Litho. Co. Wheeling. Hoffman, Bernard G. 1952. Implications of radiocarbon datings for the origin of the Dorset culture. American Antiquity, v. 18, no. 1, p. 15-17. Menasha, Wis. Jenness, Diamond 1930. A new Eskimo culture in Hudson Bay. Geographical Review, v. 15, p. 428-437. New York. Kellar, James H. 1956. An archaeological survey of Spencer County, Indiana. Indiana Historical Bureau, 68 p. Indianapolis. 242 Annals of Carnegie Museum vol. 35 Kneberg, Madeline, and T. M. N. Lewis 1955. The Archaic in western Tennessee. Mimeographed paper presented to Society for American Archaeol- ogy conference, Bloomington, Ind., 1955. Kroeber, Alfred 1944. Peruvian archeology in 1949. Viking Fund Publications in An- thropology, no. 4, New York. Larsen, Helge, and Froelich Rainey 1948. Ipiutak and the Arctic whale hunting culture. Anthropological Papers of the American Museum of Natural History, v. 42, p. 1-279. New York. Lee, Thomas E. 1954. The first Sheguiandah expedition, Manitoulin Island, Ontario. American Antiquity, v. 20, no. 2, p. 101-111. Salt Lake City. 1955a. The Archaic as seen from Ontario. Mimeographed paper presented to Society for American Archaeol- ogy conference, Bloomington, Ind., 1955. 1955b. The second Sheguiandah expedition, Manitoulin Island, On- tario. American Antiquity, v. 21, no. 1, p. 63-71. Salt Lake City. Leechman, Douglas 1943. Two new Cape Dorset sites. American Antiquity, v. 8, no. 4, p. 363-375. Menasha, Wis. Lewis, Clifford M. 1955. The Wheeling College site (460h22). The West Virginia Arche- ologist, no. 7, p. 1-8. Moundsville. Lewis, T. M. N., and Madeline Kneberg 1947. The Archaic horizon in western Tennessee. Tennessee Anthro- pology Papers, no. 2, Knoxville. Libby, Willard F. 1952. Chicago radiocarbon dates, III. Science, v. 116, p. 673-681. Washington. 1954a. Chicago radiocarbon dates, IV. Science, v. 119, p. 135-140. Washington. 1954b. Chicago radiocarbon dates, V. Science, v. 120, p. 735-742. Washington. Logan, Wilfred D. 1952. Graham Cave, an Archaic site in Montgomery County, Mis- souri. Missouri Archaeological Society, Memoir no. 2, Columbia. MacNeish, Richard S. 1952a. The archeology of the northeastern United States. In “Arche- ology of eastern United States,” p. 46-58. University of Chicago Press, Chicago. 1952b. A possible early site in Thunder Bay District, Ontario. Annual Report of the National Museum of Canada for 1950-51, Bulle- tin 126, p. 23-47. Ottawa. 1954. The Pointed Mountain site near Fort Liard, Northwest Ter- ritories, Canada. American Antiquity, v. 19, no. 3, p. 234-253. Salt Lake City. 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 243 1956. Archaeological reconnaissance of the Delta of the Mackenzie River and Yukon Coast. Annual Report of the National Mu- seum of Canada for 1954-55. p. 46-81. Ottawa. Martin, P., G. Quimby and D. Collier 1947. Indians before Columbus. University of Chicago Press. Chicago. Martin, P., and John B. Rinaldo 1951. The southwestern co-tradition, Southwestern Journal of An- thropology, v. 7, p. 215-229. Albuquerque. Maxwell, M. S. 1951. Woodland cultures of southern Illinois; Archaeological excava- tions in the Carbonclale area. Logan Museum Publications in Anthropology, Bulletin 7, Beloit College, Beloit, Wis. Mayer-Oakes, William J. 1953. An archeological survey of the proposed Shenango River Reservoir area in Ohio and Pennsylvania. Annals of the Car- negie Museum, v. 33, p. 115-124. Pittsburgh. 1955a. Prehistory of the Upper Ohio Valley. Annals of Carnegie Mu- seum, v. 34, 296 p. Pittsburgh. 1955b. Excavations at Globe Hill Shell Heap. West Virginia Archeo- logical Society, Publication Series no. 3, p. 1-32. Moundsville. McCary, Ben C. 1951. A workshop site of early man in Dinwiddie County, Virginia. American Antiquity, v. 17, no. 1, p. 9-17. Menasha, Wis. McMichael, Edward V. 1956. An analysis of McKees Rocks Mound, Allegheny County, Pennsylvania. Pennsylvania Archaeologist, v. 26, no. 3-4, p. 128-151. Milton, Pa. Miles, Suzanne W. 1951. A revaluation of the old copper industry. American Antiquity, v. 16, no. 3, p. 240-247. Salt Lake City. Miller, Rex K. 1941. McCain site, Dubois County, Indiana. Indiana Historical So- ciety, Prehistory Research Series, v. 2, no. I. p. 1-60. Indian- apolis. Morgan, Lewis H. 1877. Ancient society. 570 p. Kerr, New York. Morgan, Richard G. 1952. Outline of culture in the Ohio region. In “Archeology of eastern United States,” p. 85-98. University of Chicago Press, Chicago. Neumann, Georg K. 1952. Archeology and race in the American Indian. In “Archeology of eastern United States.” p. 13-34. University of Chicago Press, Chicago. Phillips, Philip, and Gordon Willey 1953. Method and theory in American archeology. American Anthro- pologist, v. 55, no. 5, p. 615-633. 244 Annals of Carnegie Museum vol. 35 Popham, Robert E., and J. N. Emerson 1954. Manifestations of the old copper industry in Ontario. Penn- sylvania Archaeologist, v. 24, no. 1, p. 3-19. Quimby, G. I. 1953. Indians had a lake level problem 60 centuries ago. Chicago Natural History Museum, Bulletin, v. 24, no. 6, Chicago. 1954. Cultural and natural areas before Kroeber. American Antiquity, v. 19, no. 4, p. 317-331. Salt Lake City. Ridley, Frank 1954. The Frank Bay site, Lake Nipissing, Ontario. American An- tiquity, v. 20, no. 1, p. 40-50. Salt Lake City. Ritchie, William A. 1932a. The Lamoka Lake site. Researches and Transactions of the New York State Archeological Association, v. 7, no. 4, Rochester. 1932b. The Algonkin sequence in New York. American Anthropolo- gist, v. 34, p. 406-414. Menasha, Wis. 1940. Two prehistoric village sites at Brewston, New York. Researches and Transactions of the New York State Archeological Associa- tion, v. 9, no. 1, Rochester. 1944. The Pre-Iroquoian Occupation of New York State. Rochester Museum of Arts and Sciences, Memoir 1. Rochester. 1945. An early site in Cayuga County, New York. Researches and Transactions of the New York State Archeological Association, v. 10, no. 2, p. 1-158. Rochester. 1949. The Pre-Iroquoian pottery of New York State. American An- tiquity, v. 25, no. 2, p. 97-124. Menasha. 1951a. Ground slate; Eskimo or Indian? Pennsylvania Archaeologist, v. 21, no. 3-4, p. 46-52. Milton, Pa. 1951b. Radiocarbon dates on samples from New York State. In “Radio- carbon Dating.” Memoirs of the Society for American Archaeol- ogy, no. 8, p. 31-32. Salt Lake City. 1953. A probable Paleo-Indian site in Vermont. American Antiquity, v. 18, no. 3, p. 249-258. Salt Lake City. 1955. The northeastern Archaic— a review. Mimeographed paper presented to Society for American Archaeol- ogy conference, Bloomington, Ind., 1955. Ritzenthaler, Robert E., and Paul Scholz 1946. The Osceola site, an “Old Copper” site near Potesi, Wisconsin. Wisconsin Archeologist, v. 27, no. 3, p. 53-70. Milwaukee. Ritzenthaler, Robert E., and W. L. Wittry 1952. The Ocento site— an Old Copper manifestation. Wisconsin Archeologist, v. 33, no. 4, p. 199-223. Milwaukee. Roberts, Frank H. H., Jr. 1940. Development in the problem of the North American Paleo- Indian. Smithsonian Miscellaneous Collections, v. 100, p. 51-116. Washington. 1958 Dragoo: Archaic Hunters of the Upper Ohio Valley 245 Rodgers, Edward H., and Murry H. Rodgers 1950. Archaeological investigations of the region about Lakes Mistas- sini and Albanel, Province of Quebec, 1948. American An- tiquity, v. 15, p. 322-377. Menasha, Wis. Rouse, Irving 1953a. The strategy of cultural history. In “Anthropology today,” p. 57-76. University of Chicago Press, Chicago. 1953b. On the correlation of phases. American Anthropologist, v. 57, no. 4, p. 713-722. Rowley, Graham 1940. The Dorset culture of the eastern Arctic. American Anthro- pologist, v. 42, p. 490-499. Menasha, Wis. Schmitt, Karl 1952. Archeological chronology of the Middle Atlantic States. In “Archeology of eastern United States,” p. 59-70. University of Chicago Press, Chicago. Sears, William H. 1948. What is the Archaic? American Antiquity, v. 14, no. 2, p. 122-124. Menasha, Wis. 1954. A late Archaic horizon on the Atlantic Coastal Plain. Southern Indian Studies, v. 6, p. 28-36. Chapel Hill. Smith, Arthur C. 1954. Onondaga chert in Northern Ohio. Ohio Archaeologist v. 4, no. 4, p. 9-11. Columbus. Snow, Charles E. 1948. Indian Knoll skeletons. University of Kentucky Reports in Archeology and Anthropology, v. 4, no. 3, pt. 2, p. 367-554. Lexington. Spaulding, Albert C. 1946. Northeastern archaeology and general trends in the northern forest zone. In “Man in northeastern North America.” Papers of the Robert S. Peabody Foundation for Archaeology, v. 30, p. 143-167. Andover. 1955. Prehistoric cultural development in the eastern United States. In “New interpretation of aboriginal American culture his- tory.” p. 12-27, Anthropological Society of Washington. Stewart, Paul R., and Don W. Dragoo 1954. The Gay Shriver Archaic site, Greene County, Pennsylvania. The Pennsylvania Archaeologist, v. 24, p. 106-114. Milton, Pa. Thompson A. C. 1954. The Boyer’s Run Rock Shelter— a preliminary report. Pennsyl- vania Archaeologist, v. 24, no. 3-4, p. 115-126. Milton, Pa. Webb, William S. 1946. Indian Knoll. University of Kentucky Reports in Archaeology and Anthropology, v. 4, no. 3, pt. 11, p. 115-365. Lexington. 1950. The Carlson Annis Mound. University of Kentucky Reports in Archaeology and Anthropology, v. 7 no. 5, p. 267-354. Lex- ington. 246 Annals of Carnegie Museum vol. 35 1951a. Radiocarbon dating on samples from the Southeast. In “Radio- carbon dating.” Memoirs of the Society for American Archaeol- ogy, no. 8, p. 30. Salt Lake City. 1951b. The Parrish Village site. University of Kentucky Reports in Archaeology and Anthropology, v. 7, no. 6, p. 407-451. Lex- ington. Webb, William S., and David L. Dejarnette 1942. An archaeological survey of the Pickwick Basin in adjacent portions of the states of Alabama, Mississippi, and Tennessee. Bureau of American Ethnology Bulletin 29, Washington. Webb, William S., and William G. Haag 1940. The Cypress Creek Village, site 11 and 12, McLean County, Kentucky. University of Kentucky Reports in Archaeology and Anthropology, v. 4, no. 2, Lexington. Wettlaufer, Boyd 1955. The Mortlach site in the Besant Valley of central Saskatche- wan. Department of Natural Resources, Regina, Saskatchewan. (Anthropological Series no. 1.) Willey, Gordon 1945. Horizon styles and pottery traditions in Peruvian archaeology. American Antiquity, v. 11 p. 49-56. Menasha, Wis. 1953. Archeological theories and interpretations; New World. In “Anthropology today,” p. 361-385. University of Chicago Press, Chicago. Willey, Gordon, and Philip Phillips 1955. Method and theory in American archeology, II. American Anthropologist, v. 57, no. 4, p. 723-819. Menasha, Wis. Wintenberg, W. J. * 1939. Eskimo sites of the Dorset culture in Newfoundland. American Antiquity, v. 5, p. 83-102. Menasha, *Wis. Witthoft, John 1952. A Paleo-Indian site in eastern Pennsylvania; an early hunting culture. Proceedings of the American Philosophical Society, v. 96, no. 4, p. 464-495. Lancaster. 1953. Broad spearpoints and the transitional period cultures. Penn- sylvania Archaeologist, v. 23, no. 1, p. 4-31. Milton, Pa. Wittry, Warren L. 1951. A preliminary study of the Old Copper Complex. The Wiscon- sin Archeologist, v. 32, no. 1, p. 1-18, Milwaukee. Wittry, Warren L., and Robert K. Ritzenthaler 1956. The Old Copper Complex; an Archaic manifestation in Wis- consin. American Antiquity, v. 21, no. 3, p. 244-254. Salt Lake City. (AMES L. SWAUGER AND ARTHUR M. HAYES ' t> ANTHROPOLOGICAL SERIES, NO. 4 Publication of this article has been sponsored by the Point Park Committee of the Allegheny Conference on Community Development Cover design by Clifford J. Morrow, Staff Illustrator P4-P (bE V,.*? ART. 11. HISTORIC ARCHEOLOGY AT FORT PITT, 1953 (anthropological series, no. 4) By James L. Swauger* and Arthur M. HayesJ INTRODUCTION During the period from January 15, through December 31, 1953, the Sec- tion of Man, Carnegie Museum, conducted archeological salvage work at the site of the Point State Park in Pittsburgh. Until the latter part of Febru- ary, the work was done as a part of the Section’s program under the Fund for the Study of Man, a project sponsored by the Sarah Mellon Scaife Foundation. A Service Purchase Agreement was executed between Carnegie Museum and the Department of Forests and Waters of the Commonwealth of Penn- sylvania, effective February 26, 1953, and from that time forward the Section worked as part of the Commonwealth’s general Point State Park program. Under the terms of this agreement, the Section was able to broaden and intensify its activities. A mimeographed interim report, “Archeological Salvage at the Site of Fort Pitt, 1953,” dated September 15, 1953, was prepared by the authors for the Pennsylvania Department of Forests and Waters. Copies of it were dis- tributed to interested persons. The interim report was written at that time because actual earth-moving operations at Point State Park, the phase of operation promising most likely reward for the archeologists, had ceased. We considered it desirable to inform the Department of gross results obtained as of the date of the report, since we did not believe site activity through the rest of the year would yield important results. This belief was borne out by events. This present report is based in large degree on the 1953 interim report; in fact, some of the interim report material will be found here verbatim. In a sense this present article is an expanded version of the interim report with the addition of information not present or not understood in Septem- ber 1953. PURPOSE OF THE ARCHEOLOGICAL WORK The Historical Advisory Committee of the Point Park Committee of the Allegheny Conference on Community Development recommended to the Department of Forests and Waters that, in conjunction with grading opera- tions to be conducted at the site of the Point State Park, effort be made to rescue whatever items of historical significance might come to light and be useful in development of general Point State Park plans. This recom- mendation was favorably received by the Department and was made part * Assistant Director, Carnegie Museum. tAssistant Archeologist, Section of Man, Carnegie Museum. 247 Submitted for publication, November 28, 1958 Issued June 18, 1959 248 Annals of Carnegie Museum vol. 35 of the General and Specific Conditions of the interim Point State Park contract.* Three specific duties were the responsibility of the Section of Man under this contract: 1. Collecting of all pre-1800 man-made objects. 2. Salvaging of bricks from walls that would be inaccessible later be- cause of planned highway developments. 3. Recording of gross fort features not previously discovered by archeo- logical methods, and any previously discovered for comparison with known maps and plans of the fort complex in the Point State Park area. A fourth responsibility not stated in the contract but naturally assumed by the Section was the maintenance of proper records: the taking, number- ing, listing, and initial storage of specimens and photographs; the preparing of necessary maps and drawings; the keeping of journals; the organizing of data for study and report; the keeping of strict financial records; and other normal field and laboratory routine procedures. PERSONNEL The following persons were the staff under direct supervision of the Section of Man during the Point State Park work. The specific Park duty of each is given in capital letters: James L. Swauger, archeologist, Curator, Section of Man, Carnegie Museum. Lawrence S. Thurman, historian. Curator, Old Economy, Ambridge, Pa. Arthur M. Hayes, assistant archeologist. Section of Man, Carnegie Museum. Mrs. Dorothy E. Dragoo, clerical and laboratory assistant. Section of Man, Carnegie Museum. Swauger was active throughout the whole period. Thurman worked from January 15 through May 17, when, in a letter to Swauger, he announced cessation of regular Point State Park activity because of pressure of his work at Old Economy and his poor health. Hayes worked from January 26 through June 9 when he left for summer field work with the Section’s regular field crew as agreed at the time of his hiring. Mrs. Dragoo was on part- time status from January 16 through August 19. At various times services were hired from Frank Bryan, Inc., McKees Rocks, Pa., for excavation work; Surveys, Inc., and Braun and Fulton, Pitts- burgh, Pa., for surveying work; Pittsburgh Photographic Library, University of Pittsburgh, for photographic work; Frederick R. Matson, Pennylvania State University, University Park, Pa., for ceramic study; and Carnegie Museum office staff for routine processing and final preparation of financial records and the like. * Contract for Development of Point State Park (clearing, grading, miscellaneous work), Pittsburgh, Allegheny County, Pennsylvania. Project No. 1440-1. Common- wealth of Pennsylvania, John S. Fine, Governor, Harrisburg, Pennsylvania. Depart- ment of Property and Supplies, Alan D. Reynolds, Secretary. Ralph E. Griswold & Associates, Landscape Architects, 206 Gladstone Road, Pittsburgh, Pennsylvania. Nov. 18, 1952. 1959 Swauger and Hayes: Archeology at Fort Pitt, 1953 249 Our indebtedness to numerous persons who voluntarily assisted us in one \vay or another should be acknowledged here. There were so many it would be folly to list them all, but in particular Edwin V. Pugh, Wells- ville, Ohio; Miss Rose Demorest and her staff of the Pennsylvania Room, Carnegie Library of Pittsburgh; Ralph A. McGiffin of the Department of City Planning, Pittsburgh; David W. Rial, Wilkinsburg, Pa.; and Dr. O. E. Jennings and Dr. E. R. Eller of Carnegie Museum furnished valuable in- formation. A special note of praise must be sounded for the contractor’s men and the Commonwealth’s inspectors at the Point State Park. Clauses in the contract provide for co-operation of the excavator’s people with the archeologists. The employees of the Frank Bryan company adhered not only to the letter but also to the spirit of these clauses and even exceeded them in their aid to the archeologists. C. L. Smith, Superintendent, Howard Mauk and Anthony Vice, foremen, were tireless in making certain that all curious objects and structures encountered were reported, in advising as to potential use of items of their equipment, and in suggesting ways and means to further the investigation. Not once was there any indication of impatience or lack of either interest or warm good will. They went so far as to furnish many services without charge and on their own initiative which made the arche- ologists’ lot not only more pleasant but also more useful. The same spirited interest was shown by the Commonwealth’s inspectors, John H. Reish, District Supervisor, and the resident inspectors, James K. Chambers and James K. Warren. It was a pleasure to work with these men. METHOD Time for research prior to commencement of excavations at Point State Park was limited. On January 12, three days prior to the beginning of excavation and more than a month before formal hiring, the Section was reliably informed it would be retained to perform its salvage task. Most of the Section’s preparation for its work went on while the job was in progress, but the conditions under which the Section operated were such that this simul- taneous preparation and action procedure was not a serious handicap. Most of the library research was done by Hayes. Particularly during the first three months of work there were many, many days on which the soggy condition of the site’s soil prevented digging by the contractor, days on which he could only break sidewalks or tear down stone walls. Advantage of this situation was taken by Hayes to do library work as a result of which he assembled a mass of information concerned primarily with physical features of the fort complex at the site. Both Swauger and Thurman did some library work. Miss Demorest and her staff were most helpful to us in this study. John A. Renner of Ralph E. Griswold 8c Associates supplied us with a set of maps prepared in 1943 for the Point Park Commission of the City of Pittsburgh by the Department of City Planning, with copies of the Grading Plan and Salvage and Construction Plan for the interim Point 250 Annals of Carnegie Museum vol. 35 State Park project, and with a copy of the Bliss report.* This report contains a vast amount of material concerning the fort complex. It was compiled by Wesley A. Bliss, archeologist for the Point Park Commission’s 1942-1943 investigations at the site of the Point State Park, is based largely on the archeological work done there during the three years 1941 through 1943, but also contains much other information. Without these maps and the Bliss report, we would have had no reliable guides for our own planning, work, and interpretation. The thoroughness of the Bliss report is evidenced by the fact that every check we made of his record against actual fort condi- tions as we found them proved accurate within reason. We added but little to his compilation. Rial later donated a copy of the Bliss report to the Section. Through Pugh we met McGiffin. Bliss had lodged with McGiffin while doing his work at the site of the Point State Park. McGiffin had intimate knowledge of Bliss’s work not only because of his close association with Bliss but also because much of the research on which the 1943 Department of City Planning maps were based was his. He talked to us of Bliss’s work, showed us objects recovered by Bliss, and permitted us to borrow bricks from Bliss’s pit “B” with which we could compare those we would find. Armed with the Bliss report, the City Planning maps, the interim project maps, and the McGiffin samples, we proceeded with assurance in this other- wise unfamiliar historical-archeological project. On the site we depended on observation and movement for result. Hayes worked each day from his hiring until he left. Thurman and Swauger worked alternate days until Thurman’s withdrawal from active participation, and Swauger full time from June 9 until the middle of July, and at odd inter- vals thereafter. Most of the time two men, and sometimes three, were at the site following shovels, bulldozers, high-lifts, rollers, motor cranes, and laborers as they dug into and moved the earth. Aided by known provenience of fort features and levels as determined by Bliss and shown on the City Plan- ning maps, we were able to check against the grading and salvage and con- struction plans to predict likely areas for important finds. We used the grading plan grid system as our own grid for location of finds. Although we knew most of the area from which soil was being removed (roughly the space bounded by Duquesne Way, Barbeau Street, Penn Ave- nue, and a north-south line slightly east of the Block House) was composed of fill laid down about 1900, we hoped residue of the forts and of build- ings possibly made of fort material might be found as part of this fill. This hope also kept us attentive to digging done in areas outside the known fort perimeter. To expose a portion of Fort Pitt’s brick and stone wall along Liberty Avenue (a wall now permanently covered by a ramp for the new highway development) we hired a bulldozer, pumps, and motor crane with clamshell *“ Bliss report” is the commonly used term for the unpublished “Part One of the Report of the Point Park Commission,” Pittsburgh, Pa., December 31, 1943. It was copyrighted in 1944 by the Point Park Commission. So far as the writers can dis- cover, only 12 copies of this report were prepared. The copy donated to the Section by Rial is copy 9. 1959 SWAUGER AND HAYES: ARCHEOLOGY AT FORT PlTT, 1953 251 bucket from the Bryan people. To assist in removing and transporting bricks and stones from the three pits dug here, indicated as C.M. 1, C.M. 2, and C.M. 3, we hired power hammers, trucks, and, again, a motor crane. Fig. 1 shows location of test pits (C.M. 1, C.M. 2, C.M. 3, C.M. 4 and C.M. 5) dug by Carnegie Museum, and also identifies the location of the earlier pits dug by Bliss. All of the sample removal in C.M. 1 and most of that in C.M. 3 was done by hand by Hayes and Swauger. To expose a portion of a palisade north of Penn Avenue in C.M. 4, we hired a bulldozer and high- lift for rough work, and finished off by hand. The filling of these pits was done by bulldozers. Although empowered to stop the excavator from working in any given area for fear of his damaging significant materials, we had to use this power only twice. Once north of Penn Avenue we stopped the motor crane when un- certain as to the importance of a brick structure revealed when a sewer pit was being dug. The structure proved of no value to us. Once we stopped the power shovel when we feared it might cut through the palisade in C.M. 4, and this stoppage proved wise. Fort Pitt had five bastions. Names applied to these are of relatively re- cent origin, but are accepted in literature concerning the fort. That pointing northeast is the Music Bastion; that pointing southeast, the Grenadier Bastion; that pointing south, the Flag Bastion; that pointing southwest, the Monongahela Bastion; and that pointing northwest, the Ohio Bastion. The important gross features encountered— the wall between the Flag and Grenadier Bastions exposed in C.M. 1, C.M. 2, and C.M. 3 on the south side of the fort, and the wall between the Music and Ohio Bastions, C.M. 4, on the north side, were mapped by professional surveyors. Levels on these walls were taken for us by Mauk, Grade Foreman for the Bryan company, as well as by the surveyors. We, of course, also made our own maps and drawings of these features and elements thereof. We devoted considerable attention to the making of a photographic record. All activities at the Point State Park excavation, not only of the strictly archeological work but also of the general excavation and grading procedures, were recorded in both black and white and in colored photo- graphs. The Section took 196 black and white photographs and 349 colored slides, the Pittsburgh Photographic Library took 42 black and white photo- graphs. Rial volunteered his services for the production of a group of colored slides of bricks, stones, mortar, and logs from the Fort Pitt walls, and a series of fort complex maps and diagrams in black and white transparencies. In the laboratory, activity centered on numbering, describing, and cata- loguing specimens and photographs, mounting colored slides, and pre- paring jackets for black and white photographs. Mrs. Dragoo did most of this work after initial recording of essential data by Hayes and Swauger. For recording of exact colors of Fort Pitt bricks, Mrs. Dragoo used Robert Ridgway’s “Color Standards and Color Nomenclature,’’ although for prac- ical purposes we follow Matson’s advice* and refer to bricks in their general color range as medium red, deep red, brown, etc. In the laboratory also, * Memorandum to Swauger, June 29, 1953. 252 Annals of Carnegie Museum vol. 35 Fig. 1. Location of Fort Pitt, and positions of archeological test pits 1959 SWAUGER AND HaYESI ARCHEOLOGY AT FORT PlTT, 1953 253 were finished journals, maps, and drawings necessitated by the project, and were prepared routine and special reports and letters. So far as maps are concerned, ours are mostly of detailed portions of the fort structures since the Bliss report maps are excellent and quite sufficient for general overall presentation. Swauger attended several conferences concerned with the work at Point State Park; Thurman attended some; Hayes, none. Daily journals were maintained in triplicate by each of the persons under direction of the Section. The original copy is held by the Section, the first carbon copy by the person preparing it, and the second carbon copy ac- companies the bulk of the specimens in storage. The original was held in the contractor’s office at the Point State Park for ready reference by inter- ested persons throughout the project. Forms for these journals were pre- pared by the Museum office staff and, after her hiring, Mrs. Dragoo. Forms were also prepared for entry of photographic data and expense accounts. Matson visited the site on June 29, made an examination of the bricks removed as well as those on the site, and forwarded a general memorandum to Swauger as of June 29. A representative collection of bricks and mortar was forwarded to Matson on December 22. These he will use as samples for laboratory ceramic study and as part of a collection of bricks from known sites to be assembled at the Pennsylvania State Univerity. Dr. Jennings made identifications of timbers removed from the north wall in C.M. 4, and Dr. Eller performed the same service for glass and bottles collected from the whole site. RESULTS Objects recovered Accomplishment of the Section’s first responsibility in this work in Point State Park, the rescuing of 18th century objects, was a dismal failure. So far as we can determine, we recovered not one single object except pieces of the actual fort wall which can confidently be given a pre-1800 date. We can not explain this lack of specimens. It is not due to our carelessness or lack of ability, for certainly we were diligent and equally certainly the three men who worked are capable of differentiating between late 18th century or even early 19th century objects on which some doubt might be held and late 19th or early 20th century material. We saw nothing we would hesitate to place chronologically as anywhere but in the late 19th or early 20th centuries. This conclusion as to date is readily understandable so far as most of the material moved by the Bryan company is concerned. Before we began work we knew the area had suffered flood, fire, construction disturbance, and finally a fill about 1900 covering much of it to a depth of 8 to 12 feet. But we certainly hoped to get something from work along the fort walls. The fill was disappointing. The Section had hoped to be able to accumu- late a collection from a stratified section of the fill which would give a chronological story of the city as told by its rubbish. Even this hope was dashed, for the fill was a homogeneous mass of all sorts of litter— broken 254 Annals of Carnegie Museum vol. 35 crockery, glass, sewer pipes, machine-stitched shoes, charred lumber, piles of ashes, slag, bricks, cement, tar paper, metal conduit, gas lines, etc. The size of the tools used and the immense amounts of earth displaced at one time were detrimental to locating specimens. For instance, much of the earth was torn from place by a power shovel with a bucket having a capacity of two cubic yards. The shovel took three bites to fill a truck which carried the dirt to the new fill area and dumped it. A bulldozer flat- tened the dumped earth. A sheeps’-foot roller rolled and tamped it. The process consumed about ten minutes per load and posed a pretty problem in speedy observation for the archeologists. However, the face of the shovel made a clean horizontal cut on the surface of the ground. The bucket dribbled from its mouth as it swung to the truck. The truck’s dumping made a pile with the surface easily scanned, and the edges free from packing which might hide objects. The bulldozer’s blade packed in the center of its sweep but feathered dirt out at its edges to reveal objects. The rollers churned the dump as they went bringing some things to the surface even as they tamped others down. We saw thousands of objects, and it is unlikely chance alone prevented our seeing 18th-century material. A small collection of this late fill material was made— crockery fragments, bottles, and metal objects. The work along the fort walls was begun initially with heavy machinery and was thus subject to the same handicap as work in the fill areas proper so far as observation was concerned; but all work along these walls was finished by hand and the opportunity to see was optimum. Here again we found dozens of objects, and here again we found nothing we are confident dates from the 18th century. Bliss worked many more pits than clicl we, and did them largely by hand, yet his experience with artifact recovery was no happier than ours. He was no more successful than were we in dating as 18th century any pieces re- covered that were not definitely part of the fort structure. The following statements are from the Bliss report: “The few artifacts recovered could not be definitely placed stratigraphically. Crockery, pottery, shoes, bones, fragments of iron utensils, bricks, bottles, glass, logs, stockade posts, and boat fragments made up most of the material that came out of the test pits.” p. 73. “Fragments of corckery (sic), bottles, pottery, shoes, bones, utensils, wire, nails, etc. came out of the test pits.” p. 74. With Bliss we must assume that there probably are relics of the 18th century at the site of Fort Pitt because of literary reference to the finding of such objects in the past (Bliss report, p. 75). All we are certain of is that we found none. The following categories and numbers of objects were recovered and catalogued: Bricks and brick fragments 357 Mortar and boxes of crushed mortar 53 Stone from fort walls 19 Boxes of rubble 4 Bone 1 1959 SWAUGER AND HAYES: ARCHEOLOGY AT FORT PlTT, 1953 255 Oyster shell 1 Metal fragments 5 Pottery and crockery fragments 20 Bottles and glass fragments 54 Stone paving blocks 5 Logs and bark and bark fragments 14 Cement block 1 Marble slab 1 Metal ball from Mt. Washington 1 Motorcycle license 1 Unidentified 4 Total 541 Of the 541 objects catalogued, 466 numbered items were taken to the building of the Historical Society of Western Pennsylvania at 4338 Bigelow Boulevard, Pittsburgh, on December 16, 1953. They remain there until the construction of a museum building at Point State Park permits their display in their area of origin, or until other use is found for them by the Historical Advisory Committee. The other 75 items were retained by Carnegie Museum as Accession 15661, in its permanent collection as a representative series of objects from the 1953 work at Fort Pitt. Of these 75 items 22 were forwarded to Matson on December 22, for his study. Six items were given as tokens of appreciation to persons instrumental in the Fort Pitt operations. Record of the distribution of numbered items is in the catalogue book. Brick salvage We were more successful in the task of salvaging bricks. Pits C.M. 1, C.M. 2, and C.M. 3 were dug specifically to locate the fort walls to be covered by a highway ramp paralleling Liberty Avenue to the north and to remove from them all bricks that could be saved. As will be explained in more detail in discussion of fort features, brick wall was found in C.M. 1 and C.M. 3; none in C.M. 2. All bricks taken from C.M. 1 and a few of those from C.M. 3 were retained as samples. Most of the bricks from C.M. 3 were salvaged for later usage at Point State Park. Initially we removed bricks by hand with hammers, chisels, and bars. The bricks are so soft however, the mortar so hard, relatively, that the pounding necessary to separate bricks from mortar shattered so many bricks we estimate a loss of 50% in C.M. 1 from this process. Finding the same situation recurring in C.M. 3, we decided to use power tools, deliberately losing some rows of bricks in order to get out great chunks of brick wall. This process was far more successful than the first tried, and we estimate a loss not exceeding 30%. Several hundred good bricks in great blocks and many loose ones were removed by use of power hammers, lifted by motor crane, taken by truck to the northwest comer of the intersection of Penn Avenue and Barbeau Street, and there buried by motor crane. We salvaged by no means as many bricks as were desired to carry out initial restoration plans for the Monongahela and Ohio Bastions as out- lined for us at the beginning of the project. This is due in part to the 256 Annals of Carnegie Museum vol. 35 difficulty of pulling good whole bricks from their seats in the walls, but mostly to the fact that there are not nearly so many bricks left in the walls as had been hoped. Salvage operations by 18th century contractors were evidently very thorough,* and they didn’t leave many bricks for us to re- cover. Those bricks we did salvage, however, will aid in the manufacture of new bricks on the old model for the restoration work. Some mention must be made of the bane of the archeologists’ existence- souvenir hunters. From beginning to end of the excavation we were badgered by spectators for pieces of brick and stone, mortar and wood, from the fort walls. We did give away some valueless fragments of authentic material and some whole bricks, chiefly to members of the Bryan company and other persons who were working with us on the site. We also gave away, and people took, modern bricks from the fill, fragments of cement from the railroad retaining walls, chunks of slag and cinders, bags of earth from the fill, and the like. Some of the more persistent souvenir seekers were ap- parently quite satisfied with bricks from the sidewalk on Liberty Avenue, bricks stamped plainly “Toronto” and ‘‘Pittsburgh Buffalo Co.”, which they evidently mistook for Fort Pitt bricks. For future reference, only those objects carrying the catalogue designation “PPP/number” on a square of yellow paint, or those authenticated by a certificate signed by Hayes or Swauger should be considered actual Fort Pitt material. Fort features General. Gross features of the fort were located in the four pits dug at the request of the Section and numbered C.M. 1, C.M. 2, C.M. 3, and C.M. 4. It is likely part of the wall between the Flag and Grenadier Bastions was found in a hole dug to cap a water line under Liberty Avenue. This pit we numbered C.M. 5. We did no work with C.M. 5 except to collect bricks as samples and to photograph the wall since it was inexpedient to close Liberty Avenue to traffic. Fig. 1 gives the positions of these features. Test pit C.M. 1. In this pit, 38.35 feet of the fort wall were uncovered. This portion of the fort’s eastern wall is the angle at the southwest corner of the Grenadier Bastion and is composed of part of the southwest arm of that bastion (We call this the eastern arm of the angle) and part of the arm running more southerly into Liberty Avenue where it angles more or less east into the Flag Bastion (We call this the western arm of the angle). These arms are illustrated in Fig. 2. Both arms of the angle are interrupted by a cement wall 14.43 feet from the corner along the eastern arm, 23.92 feet along the western arm. Only part of the original wall exists. In the corner of the angle were perhaps two hundred whole bricks and many fragments. Most of the whole bricks were removed for samples, those directly at the corner being left for the benefit of future diggers. A section was taken through the stone body of the wall to determine its form (Fig. 6) and a profile cut was made at the foundation’s southern face to determine structure and depth. Stones from the wall were removed for samples. Mortar samples were taken from both the brick facing and the stone wall. * Pennsylvania Archives, 1854, v. 10, p. 483. 1959 SWAUGER AND HAYES! ARCHEOLOGY AT FORT PlTT, 1953 257 CORNER OF ANGLE TQ mark on curb- 44. o' EL. FT., MIDWAY EASTERN ARM-8' EL. PT., MID WAY WESTERN A RM - 12* MARK TO MON., NE CORNER, INTERSECTION OF PENN AND BARBEAU-495^ jjG- 1 ABOUT !"■ 20' SKETCH SCALE 1 GENERAL PLAN PROFILE AREA DIRECT COPY, FIELD SKETCH OF 15 APR., *53 Fig. 2. Location of test pit C.M. 1 258 Annals of Carnegie Museum vol. 35 The wall averages 8 to 10 feet in width. It is made up of four “steps”. The lowest of these we call “foundation”; the second, “footer”; the third and fourth, “wall”. At the profile section the height of structure remaining is 4.92 feet, of which 2.92 feet is foundation; 0.5 foot, footer; 1.55 feet, wall. Levels at the footer of the western arm average 718.67 feet above the sea-level datum plane, comparing with Bliss’s levels averaging 718.4 feet for this part of the fort wall.* The footer of the eastern arm is at an average elevation of 717.52 feet, it being about one foot lower than the footer of the western arm, and this compares favorably with the Bliss level given above. The high- est elevation for the portion of wall remaining in C.M. 1 is 720.17 feet. The foundation is made of eight layers of flat, roughly face-dressed stones averaging 1.5 feet in width, and a little more than 0.25 foot in thickness, with the exception of the top course which is carefully face-dressed stone, with an average thickness of 0.33 foot. Some of these foundation stones are mortared to each other; others are not. All had been laid with care and are quite plumb. Excavations at the rear of the profile cut were not carried to the base but revealed a vertical wall corresponding to that in the front but much more carelessly laid, and not mortared. A ditch, commonly called the moat, ran along the eastern side of the fort. Bliss located it in his first test pit, “A,” and established its elevation. In the moat area, measured vertically from the top course of the founda- tion, we have 1.83 feet of clay and fine gravel to a layer of heavy gravel and black muck which continues 1.42 feet to the base of the foundation. This black muck and gravel layer has a superior elevation of 716.67 feet and compares favorably with Bliss’s elevation of 716.6 feet for the floor of the moat. The general relationships of this and other features of the wall can be best understood by reference to Fig. 2, where a copy of the field sketch of the profile is shown. The footer is set back an average of 0.67 foot from the face of the foundation, consists of two courses, one a shimming layer averaging 0.17 foot thick, the other the footer stones proper averaging 0.33 foot thick. Each of these courses is carefully face dressed, and the footer stones were laid with such precision that it is impossible to pry one loose without levering out its neighbors to right and left at the same time. The footer facing stones average 1.5 feet in length, and maintain their 0.33-foot thick- ness for 0.75 foot back from their faces to where, from a definite transverse ridge, they taper from the top back to an average thickness of 0.22 foot, the bottoms remaining level. The footer stones proper average one foot in width. Since the wall proper is set back almost two feet from the face of the footer, thin stones laid along the sloping rear of the footer stones carry their line back into the wall where it is lost in the stones of the wall. The wall at the profile cut is made up of two steps. The first is set back 1.92 feet from the footer face, and averages 0.33 foot in thickness. It is a single layer of carefully face-dressed stones. The second is set back 2.75 feet from the footer face, and averages 1.25 feet in thickness. It is made up of four layers of roughly face-dressed stones. Most of the wall stones average a little * Bliss report, p. 76. 1959 SWAUGER AND HAYES: ARCHEOLOGY AT FORT PlTT, 1953 259 more than one foot in width, and are roughly rectangular in most instances, but there is great variation in shape of those which are not rectangular. These eccentric stones were undoubtedly used to fill gaps in the more regular courses where errors of judgment as to size on the part of the masons had to be rectified. Corresponding to the superior elevation, of the foundation and 6.42 feet back from its face is a peculiar set of two “steps” completely incased in other wall stones. The lower of these is 0.83 foot thick and is made up of four courses of stone. The higher is 0.46 foot thick and is composed of two layers of stones. All stones used average 0.75 foot in width. All are care- fully face dressed and are tightly mortared together. The presence of this subsidiary and “hidden” wail suggests that after the foundation had been laid a mortared wall was built up corresponding to the front vertical wall face and more than half way to the rear face. We can conjecture no specific purpose behind this method of construction. Most of the wall stones are carefully mortared for two rows back of the interior wall discussed above. The final two or three courses, however, are carelessly set in mud packing. This is as true of the highest remaining layers of the wall as of the rear of the footer and foundation. We were impressed by the exceeding hardness of the mortar used and by the great quantity used. In many instances it was easier to chisel stone away from mortar than mortar away from stone while the profile cut was being made. In many areas of the profile cut, we found the mortar was thicker than the stones it held together. Almost it looks as if the masons had poured great batches of mortar over a layer of stones and set the next course on the mortar to sink to position by weight alone. The bricks were used as facing for the stone wall. In test pit C.M. 1 the bricks had been worked at so much in times past that it is now difficult to determine the number of bricks used in each course either horizontally or vertically. In test pit C.M. 3 the number was more readily apparent and a full discussion of the bricks will be given in the section devoted to that pit. A total of 111 bricks and brick fragments, 50 mortar samples, 18 wall stone samples, and some miscellaneous boxes of rubble, bone, shell, glass, and metal pieces were removed from C.M. 1. We can not say that anything but the bricks, mortar, and stones are 18th century. Test pit CM. 2. Less than 10 feet of the arm running south into Liberty Avenue were uncovered. The position is shown in Fig. 3. It is only foundation, footer, and very little wall, all corresponding to those elements in C.M. 1. No bricks are present. Since at the time C.M. 2 was opened we were search- ing primarily for bricks, the portion here had no bricks and could tell us nothing we had not learned in C.M. 1, and by contract we were expected to get on with our work as rapidly as possible, we did no work at this portion of the wall once its general character had been ascertained. This decision was productive of good results, for Smith, with Renner’s approval, diverted into C.M. 2 a sewer line planned to run directly through 260 Annals of Carnegie Museum vol. 35 Fig. 3. Location of test pits C.M. 1, C.M. 2 and C.M. 3 1959 SWAUGER AND HAYES: ARCHEOLOGY AT FORT PlTT, 1953 261 Fig. 4. Location of test pit C.M. 4 262 Annals of Carnegie Museum vol. 35 C.M. 3 at almost footer depth, and we were thus free to work in C.M. 3 without interference. Test pit C.M. 3. In this pit, 33.92 feet of the arm coming south from C.M. 1 and C.M. 2 were uncovered. The position is shown in Fig. 3. The ex- cavation for this pit ran into Bliss’s pit “B” and included about ten feet of the wall he found. Fig. 5. Portion of wall of Fort Pitt between Grenadier and Flag Bastions. Test pit C.M. 1 Here is a section of wall consisting of 1.83 feet of foundation with a top elevation of 718.37 feet, comparing favorably with Bliss’s elevation for this element,* and a base elevation of 716.54 feet where it is bedded in gravel below a layer of muck. Above the footer, which is of the same character as in C.M. 1, is a brick wall laid in English bond, 2.17 feet high with a backward slope of 1.02 inches in this height. There is then a stone wall 2.67 feet high above the brick wall. From the face of the foundation to the rear of the stone wall the average distance is 9.67 feet. Fig. 7 gives a good general view of test pit C.M. 3. The stone wall is built in a series of four steps above the footer. Bricks are in place on the footer, as shown in Fig. 8, and for a height of a little * Bliss report, p. 76. 1959 Swauger and Hayes: Archeology at Fort Pitt, 1953 263 over two feet above it, but it is probable they once faced the whole wall. The bricks are only four layers wide on the footer but due to their “reverse step” method of laying, they run eight to ten layers wide two feet above it. They tie into the stone wall not only by means of the “reverse steps,” but also by heavy mortar bedding. In one or two places, stones for the wall above a layer of bricks extend over the layer and are mortared to them. This is a random occurence, however. Fig. 6. Section through wall of Fort Pitt between Grenadier and Flag Bastions. Test pit C.M. 1 Bricks average 8.5 inches long, 3.5 inches wide, and 2.25 inches thick. Exceptions to this standard occur, particularly among those bricks resting directly on the footer, and usually the eccentricity is in length, some bricks being as much as 1 1 inches long. A few bricks, too, are not truly rectangular on all faces but are chopped and hewed to different angles and curves to fit special niches in the walls where bricks of standard size and shape were not efficient. A representative two-square-foot section of the wall shows in its face 12 stretchers and 24 headers. Fig. 9 gives a good view of the brick wall. The bricks, and here we derive most of our information from Matson’s memorandum of June 29, are of a red burning clay with a light intermixture of river sand or a clay rich in fine-grained sand. Some of the brick are dense, heavy for their size, others almot punky in texture, light for their 264 Annals of Carnegie Museum vol. 35 size. Probably these discrepancies are due to positions of bricks in the kiln. In color they fall generally into the deep and medium red classes (We here follow Matson in describing in a few general visual categories rather than in the specific categories used by Mrs. Dragoo in preparing her catalogue descriptions) with some pale orange, and some brown bricks. Some are burned black around the edges. Some have vitreous glazing over portions of them. Fig. 7. Brick facing on wall of Fort Pitt between Grenadier and Flag Bastions. Test pit C.M. 3 The bricks were made in a mold, probably one of wood. This mold formed three sides and both ends of the bricks, the upper broad surface being formed by tamping with fingers or palms of hand. There usually is a little overhang on the upper surfaces, and in broken bricks, the irregular upper surfaces are clearly seen. The mortar is light brown, a clayey, sometimes muddy looking mixture, with considerable lime washing and lumps included. Some brick lumps are found in the mortar but Matson does not believe they were included as grog. There is a great deal of variation in the amount of mortar used in any two groups of bricks, and an even greater variation among groups of stones in the wall. 1959 SWAUGER AND HAYES I ARCHEOLOGY AT FORT PlTT, 1953 265 Construction of the wall, details of the foundation and footer, and stratig- raphy in the moat area are the same as in C.M. 1. The situation in regard to specimens is also the same as that in C.M. 1; that is, only objects pulled from the wall can be considered 18th century material. Altogether, 216 bricks and brick fragments, two mortar samples, Fig. 8. Detail of brick facing at footer of wall of Fort Pitt between Grenadier and Flag Bastions. Test pit C.M. 3 266 Annals of Carnegie Museum vol. 35 and one stone sample were removed from C.M. 3 and kept as specimens. Many other bricks and some stone, as noted earlier, were buried. Test pit C.M. 4. As illustrated in Fig. 5, 175 feet of a stockade line paral- leling the wall between the Music and Ohio Bastions were exposed. The po- sition of this line as surveyed by Braun & Fulton does not correspond exactly to features given in Sheet A2 of 7, “Forts and Fort Features . . .”, of the Department of City Planning maps. It is our opinion that it represents the Fig. 9. Detail of brick facing of wall of Fort Pitt between Grenadier and Flag Bastions. Test pit C.M. 3 interior line of the earth wall structure between the Music and Ohio Bastions. The portion of the fort revealed by excavation at C.M. 4 consists chiefly of a stockade line 147 feet long running almost due east and west, its eastern end lying 356.26 feet west of the monument at the northeast corner of the intersection of Penn Avenue with Barbeau Street, and 101.94 north of the 7-foot running line on the north Penn Avenue sidewalk. Its western end is 494.8 feet west of the monument, 146.06 feet north of the running line. It is interrupted in several places by gas lines, by intrusive buildings, and by blank areas which have suffered disturbance but of exactly what type we did not determine. In general it is whole, and, as illustrated in Fig. 10, it is quite plain. 1959 SWAUGER AND HAYES! ARCHEOLOGY AT FORT PlTT, 1953 267 At the eastern end of the main part of the stockade, 15.6 feet of stockade incline almost due north. It is a clearly defined section but is interrupted by a fill at its northern end. We excavated across the line of its direction north of its last post, but although we went to a cement railroad retaining wall, we had no success. At the western end of the main part of the stockade, about 12.5 feet of post line is barely discernible inclining to the south. We traced it by Fig. 10. Top of line of stockade posts, north wall of Fort Pitt. Test pit C.M. 4 means of organic smudges, remnants of posts, as illustrated in Fig. 11, and one fairly well preserved post top, this latter the most southerly element found. Excavation out its line of departure met with no success. Elevations taken for us by Mauk average 720 for the tops of the stockade posts. We doubt that these levels are significant so far as the wall itself is concerned, but they may be in establishing the height of the parade ground, for obviously the posts had been chopped off almost at ground level at some time in the past probably so buildings could be constructed with either foundations or even first floors at this 720-foot level. The stockade, as shown in Fig. 12, is composed of two types of posts, large ones averaging 0.83 to 1 foot in diameter, with six feet of their length well preserved and in place, and smaller ones averaging 0.33 to 0.5 foot in 268 Annals of Carnegie Museum vol. 35 Fig. 11. Smudges and remnants of west portion of line of stockade posts, north wall of Fort Pitt. Test pit C.M. 4 diameter, with an average length of three feet well preserved and in place. The large posts had not been sharpened, had not been driven into position in the wall by mauls, but had been stood in a ditch dug to receive them as indicated in Fig. 4. Earth had then been tamped in around them. At the one vertical section cut, a short horizontal log was still in place lying snug against the northern face of the post line and covered with tamped earth. 1959 SWAUGER AND 1 1 AYES: ARCHEOLOGY AT FORT Pi I T, 1953 269 The shorter posts had been sharpened and driven into the filled ditch to strenghen the post line and to fill interstices between the large posts. While it is difficult to to be certain of the number of posts remaining, since the entire stockade line was not completely cleaned and profiled, 159 large, and small post tops were counted in the wall, and it is likely the exact total is not significantly different. Logs brought to the museum as samples were identified by Dr. Jennings as elm and black oak. Fig. 12. Profile cut revealing stockade posts in place, north wall of Fort Pitt. Test pit C.M. 4 The superior level of a representative post at the vertical section is 723.47 feet. Subtracting the post’s length, six feet, from this figure, gives a level for the base of this particular post, and essentially a practical level for the posts in the section, of 717.47 feet, very nearly the elevation for the footing of the stone wall.# This is, of course, slim evidence, but it may in- dicate an effort to have the bases of all walls at nearly the same level. East of the stockade line and to all practical purposes on line with it at a distance of 10 feet is a burned stump chopped off at a level corresponding to the tops of the stockade posts; beyond that 3.83 feet a square-cut post, that we do not believe is contemporary with the fort, chopped off at the same level; and beyond that 8.17 feet another burned stump chopped off as was * Bliss report, p. 76. 270 Annals of Carnegie Museum vol. 35 the first. Ten feet beyond this stump, 32 feet from the stockade line, and still on line with it, were found three sharpened stakes which were recovered. These average a little over three feet in length, and, as identified by Dr. Jennings, are hickory, red oak, and beech. It has been conjectured they were cheveaux de frise, but since they are not robust, they may have been impaling stakes for sinking in the moat with their points up, or they may have been only stakes used in stretching lines. From test pit C.M. 4, six bricks and brick fragments, three pieces of crockery, and 13 pieces of wood and bark were removed as samples. The wood and bark we are certain are 18th century. The bricks and crockery we feel sure are 19th or early 20th. Test pit C.M. 5 was not dug at the request of the Section of Man but is numbered and mapped. Its potentials were discussed, earlier, and its position is shown in Fig. 1. The pump. The map Sheet A2 of 7 mentioned on page 266 shows a pump about sixty feet south of the west end of the stockade line. At approxi- mately this position normal grading operations uncovered a tub-like struc- ture with its exterior made of vertical staves like those of a barrel, and its interior made of alternating circles of bricks and curved pieces of wood. Digging in this structure revealed only a mass of fine, dark dust shot with silvery specks of metal. The bricks were inspected by Matson during his visit on June 29, and he agreed with our conclusion that they do not resemble known fort bricks. The contractor was told the element was not of con- suming interest to us, and it was covered as part of the regular grading work. Since that time, however, we have seen an illustration* of a 17th century well barrel at Jamestown, Va., which to some extent resembles the structure described above except for the interior lining. It is not impossible that this may have been part of the catch basin for the pump indicated on the “Forts and Fort Features . . .” plan although it is not probable. RECORDS Note-books, journals, correspondence, etc. Considered as an integral part of the collection of material made during work at the Point State Park are the records maintained. These are briefly discussed here with notice as to their locations, copies, and the like, for the benefit of future workers at the Fort Pitt site. Journals were kept by Hayes, Thurman, and Swauger. The books them- selves remain in the possession of those who made them, but typed copies are held by the Section, as well as a copy of the journal prepared by Mrs. Dragoo. Duplicate copies of these journals accompany the specimens at present in the building of the Historical Society of Western Pennsylvania. They will move with the objects to the Point State Park museum. In such records, of course, the Bliss report must be given a place. It is history of the Point area, of forts such as was Fort Pitt, specific discussion of specific items, and discussion of specific archeological activities. *Fig. 195, illustration for Jean C. Harrington, "Historic Site Archeology in the United States.” In "Archeology of Eastern United States,” James B. Griffin, ed., University of Chicago Press, Chicago, 1952. 1959 SWAUGER AND HAYES: ARCHEOLOGY AT FORT PlTT, 1953 271 Swauger’s field note-book, a mining transit book, remains in the Section’s possession. It contains only a few items of textual information not con- tained in the journal but does contain sketches which may be of future use including, for instance, one showing the location of the burial pit for bricks and stones from C.M. 3. Copies of all reports made to the director of the museum or to various officers of the Plistorical Committee are kept by the Section. These include progress reports, estimates of the situation as the excavations developed, and the like. The Matson memorandum of June 29 is held with these reports. A detailed descriptive catalogue of the objects recovered, giving assigned number, origin within the excavated area, full description, and disposition of each object was drawn up by Hayes, Mrs. Dragoo, and Swauger. The original is in two volumes. Typed copies were also made. One of these is held by the Section; one accompanies the specimens. Copies of correspondence are kept in the Point State Park record books by the Section. Used in concert with the journals, these copies provide an excellent listing of persons concerned with the project either officially or with only a general interest. The Section has kept a fairly full publicity record of its activities at the Point State Park. At present the clippings which compose this record are kept in the Point State Park record books but in time they will be transferred to the Section clipping file where they will always be available for study. Maps and drawings The following groups of maps and drawings are held by the Section as part of the Point State Park records. The first seven are those prepared by the Department of City Planning of the City of Pittsburgh for the city’s Point Park Commission in 1943. Their scale is 1"=50'. 1. Archaeological Excavations of 1942-1943, Location of Test Pits and Fort Features. Sheet A-l of 7. 2. Forts and Fort Features of 1754 to 1792 Superimposed upon a 1942 Map of Point Area. Sheet A-2 of 7. 3. Undisturbed Sub-surface Relevant to Surface and Sub-surface Fort Features Showing Basement Areas. Sheet A-3 of 7. 4. Physiographic Changes from 1754 to 1942 Relevant to Surface and Sub-surface Fort Features Superimposed upon a 1942 Map of Point Area. Sheet A-4 of 7. 5. Sub-surface Utilities Relevant to Surface and Sub-surface Fort Features Superimposed upon a 1942 Map of Point Area. Sheet A-5 of 7. 6. Owners of Record and Character of Buildings as of 1940. Sheet A-6 of 7. 7. Land and Buildings Assessed Values as of 1940. Sheet A-7 of 7. The next four maps were prepared by Ralph E. Griswold & Associates in 1952. 8. Plot Plan, Point State Park. Scale l"=50r. 9. Survey Diagram, Point Park. Scale 1"=100'. 10. Grading Plan, Point State Park. Scale V' — bO'. 11. Salvage and Construction Plan, Point State Park. Scale 1"=50'. 272 Annals of Carnegie Museum vol. 35 The surveyors hired by the Section produced these maps: 12. Carnegie Museum Pit 1, Point Park project. Rudolph Agresti. March. 1953. Scale l"-5'. 13. Uncovered Portion of Stockade of Original Fort Pitt in Point Park. Braun and Fulton. July, 1953. Scale 1" = 50'. Swauger prepared two drawings labeled “Sketch Maps” on tracing paper for overlay use during the excavation work. Their scale is 1"=50'. They are: 14. Sketch Map no. 1, Excavations and Results, Bliss Test Pits. February 21. 1953. 15. Sketch Map no. 2, Probable Excavation Area, P.P.P. February 22, 1953. Swauger and Flayes prepared a series of drawings labeled “Charts” for use in both planning and record procedures. Their scales vary from chart to chart and in instances within charts for either horizontal or vertical read- ings. In such instances, the first scale given is horizontal, the second, vertical. In the listing given here, Swauger’s charts are identified by the initials “J.L.S.”; Hayes’s by “A.M.H.”. 16. Chart 1. Point Park Elevation Chart, No. 1: “O” line. Feb. 4., 1953. J.L.S. Scale 1"=50'; 0.2"= 1'. 17. Chart 2. Point Park Elevation Chart, No. 2: “N-l” line. Feb. 4, 1953. J.L.S. Scale 1"=50'; 0.2"= T. 18. Chart 3. Point Park Elevation Chart, No. 3: “N-2” line. Feb. 7, 1953. J.L.S. Scale 1"=50'; 0.2"= T. 19. Chart 4. Point Park Elevation Chart, No. 4: “N-3” line. Feb. 7, 1953. J.L.S. Scale 1"=50'; 0.2"= 1'. 20. Chart 5. Point Park Elevation Chart, No. 5: “N-4” line. Feb. 9, 1953. J.L.S. Scale 1"=50'; 0.2"= T. 21. Chart 6. Point Park Elevation Chart, No. 6: “S-l” line. Feb. 9, 1953. J.L.S. Scale 1"=50'; 0.2"= T. 22. Chart 7. Point Park Elevation Chart, No. 7: “S-2” line. Feb. 9, 1953. J.L.S. Scale 1"=50'; 0.2"= T. 23. Chart 8. Point Park Elevation Chart, No. 8: “S-3” line. Feb. 9, 1953. J.L.S. Scale 1" = 50'; 0.2"= T. 24. Chart 9. Point Park Elevation Chart, No. 9: “S-4” line. Feb. 9, 1953. J.L.S. Scale 1" = 50'; 0.2"= V. 25. Chart 10. Stratigraphic Charts, Wesley Bliss test pits. Feb. 21, 1953. J.L.S. No horizontal scale; vertical scale, 0.2"= T. 26. Chart 11. Stratification, Bliss test pits, Land Wall. Feb. 22, 1953. J.L.S. Scale 1"=50'; 0.2"= 1'. 27. Chart 12. Stratigraphic Charts, Contours and Pits, S6-S7 Contours, Pits “A” and “B.” Feb. 24, 1953. J.L.S. Scale 1"=50'; 0.2"= T. 27. Chart 13. Stratification and Horizontal Plan, Land Wall. Feb. 24, 1953. J.L.S. Scale 1"=50'; 0.2"= T. 28. Chart 14. Stratification in Pits, Southeast Wall. March 14, 1953. J.L.S. Scale 1"=50'; 0.2"= T. 29. Chart 15. Pit Stratification, C.M. 1, Bliss pits “A” and “B.” March 22, 1953. J.L.S. No horizontal scale; vertical scale, 1"=1'. 1959 SWAUGER AND HAYES! ARCHEOLOGY AT FORT PlTT, 1953 273 30. Chart 16. Archeological Knowledge, Southeast and East Walls of Fort Pitt. (Prepared for presentation with report of April 7, 1953, to Gris- wold from Swauger.) April 7, 1953. J.L.S. Scale 1"=50'; 0.2"= T. 31. Chart 17. Point Park Project, Miscellaneous Figures, Carnegie Mu- seum pit 1. March 22, 1953. J.L.S. Varying scales. 32. Chart 18. Point Park Project, Miscellaneous Figures, Carnegie Mu- seum pit 1. March 22, 1953. J.L.S. Varying scales. 33. Chart 19. First, second and third sections, Retaining wall, C.M. 1. March 25, 1953. A.M.H. 0.1"= 1". 34. Chart 20. Various views, Wall in C.M. 3. A.M.H. April 10, 1953. 1"=2'. 35. Chart 21. Fourth and fifth sections, Retaining wall, C.M. 1, April, 1953. A.M.H. 1"= V. 36. Chart 22. Various views, Wall in C.M. 3. A.M.H. April 22 to May 1, 1953. A.M.H. Varying scales. Photographic records There are three categories in the photographic record: 1. Kodachromes taken chiefly by Swauger and Rial. 2. Black and white photographs taken chiefly by Swauger. 3. Black and white photographs taken by the Pittsburgh Photographic Library. Kodachromes, negatives and prints of Swauger’s black and white photo- graphs are in the possession of the Section, held as part of the Point State Park record, and are entered and numbered in routine fashion in the Section’s photographic file. Prints of the Pittsburgh Photographic Library’s black and white photographs are held by the Section; the negatives are re- tained by the Pittsburgh Photographic Library, where prints can be obtained. The prints from the Pittsburgh Photographic Library bear that Library’s numbers. Not all the photographs taken by Swauger (and his pictures form the bulk of the photographic record) are good photographs. However many that are blurred in places through being out of focus or because of movement are retained because portions of the slide or print retain sharp impressions of part of the work. They are part of the listing below which gives the categories of subjects, the kind of picture, the taking agency, and the proper reference numbers in the files of the taking agencies. A complete descriptive list, slide by slide, print by print, that includes an estimate of the condition of the slide as a photograph is held by the Section. 274 Annals of Carnegie Museum vol. 35 List of photographs Kodachromes Black and white Black and white Swauger and Rial Swauger Pittsburgh Photographic Library General views of the site 901-905, 908, 911, 678, 679, 686-690, 912, 918, 919, 923- 746, 747, 818-822, 936. 939-958, 1561, 829-834. 1567-1573, 1576- 1578, 1591-1602, 1605, 1612, 1619, 1656, 1661, 1663- 1681. Historic markers, buildings, etc. 913-917,920, 921, 937, 938, 1581-1590. General use of heavy equipment 906, 907, 909, 910, 922, 951-954, 1532- 1537, 1562-1566, 1574, 1575, 1579, 1580, 1603, 1604, 1606-1611, 1613- 1618, 1652, 1655, 1657-1660, 1662, 1678. Test 959-999, 1500-1506, 1508-1513, 1524-1531, 1540. 680-685, 691, 692, 770, 789, 807-817, 823-828, 835-837, 868-874. pit C.M. 1. 700-745, 748-764, 769, 771-774, 778- 781. 11608-11611. 11600-11607, 11621- 11623. 1518. 1507, 1514-1517, 1519-1523, 1538, 1539, 1541-1560, 1643. 1620-1642, 1644- 1651, 1653. Test pit C.M. 2. Test pit C.M. 3. 765-768, 776, 777, 782-788, 790-806, 854. Test pit C.M. 4. 838-853, 855-867. 1683-1750. Test pit C.M. 5. 693-697. 11593-11599, 11612- 11620. 11566, 11568-11580. Miscellaneous 1654, 1682. 698, 699. 11577. ART. 12. SYSTEMATIC NOTES ON NORTH AMERICAN BIRDS 3. The Northeastern Races of the Long-billed Marsh Wren ( Telmatodytes palustris) By Kenneth C. Parkes Associate Curator of Birds, Carnegie Museum This is the third of a series of papers on the systematics and nomenclature of certain North American birds. A general introduction and acknowledg- ments for the series as a whole will be found in the first paper (Parkes, 1955). In connection with the present paper I wish to add my thanks to Mr. Robie W. Tufts of Wolfville, Nova Scotia, through whose good offices I was able to examine two of the three known Nova Scotia specimens of Long-billed Marsh Wren (the third, in the National Museum of Canada, I have also seen). Most of the work on which the present paper is based was done in 1949. The Long-billed Marsh Wren ( Telmatodytes palustris) has a wide distribu- tion in North America. Like the Sharp-tailed Sparrow (Ammospiza caud- acuta), it has both inland, fresh-water populations, and populations con- fined to narrow coastal strips. Unlike the Sharp-tailed Sparrow, the Marsh Wren’s range reaches the Pacific coast as well as the Atlantic. This study is restricted to the tangled taxonomic and nomenclatorial history of the Long- billed Marsh Wrens in that portion of the species’ range lying east of the Rocky Mountains and north of the mouth of Chesapeake Bay. As of the fourth edition of the A.O.U. Check-list (1931), the birds of the area in question were divided as follows: Telmatodytes palustris palustris (Wilson), Long-billed Marsh Wren. Telmatodytes palustris laingi Harper, Alberta Marsh Wren. Telmatodytes palustris dissaeptus (Bangs), Prairie Marsh Wren. These vernacular names will be frequently used for convenience in the dis- cussion beyond, since the usage of the subspecific names has been subjected to much shifting back and forth. Todd (1937) claimed that Bangs’ dissaeptus (type from Wayland, Massa- chusetts) was not available for the Prairie Marsh Wren, as he (Todd) was unable to separate Bangs’ type series from a series Todd considered to repre- sent typical palustris. The A.O.U. Check-list Committee apparently agreed with Todd, and in the Twentieth Check-list Supplement (Wetmore et al., 1945, p. 446) the name for the Prairie Marsh Wren reverted to the next oldest available name, iliacus Ridgway (type from Wheatland, Indiana). Todd’s 1937 paper was cited as authority for this change. Todd had mentioned certain differences between the type specimen of iliacus and other specimens from neighboring localities, but attributed these differences to possible post-mortem color changes. Aldrich (1946, p. 131) claimed that the type specimen of iliacus represented a migrant of the Alberta subspecies known to that time as laingi Harper. Since the A.O.U. Check-list Committee agreed with Todd in 1945 that dissaeptus was a synonym of palustris , and (assuming Aldrich to have been correct about the identity of the type of iliacus) since the name iliacus applied properly to the Alberta Marsh Wren, it would appear that the Prairie Marsh Wrtn was without a name. Oddly enough, the A.O.U. Check-list Committee reversed 275 INSTITUTION J-0: 1 3 fSfft Submitted for publication, March 24, 1959 Issued, July 10, 1959 276 Annals of Carnegie Museum vol. 35 itself in the Twenty-third Supplement (Wetmore et aL , 1948, p. 441), citing Aldrich as authority for a change from iliacus back to dissaeptus for the Prairie Marsh Wren. Actually, Aldrich left the question open, confining himself merely to a reidentification of the type of iliacus as a side remark in the description of a new race from the far West. He made no mention of either the name dissaeptus or the population to which it had been applied. The A.O.U. committee gave no indication as to why they no longer agreed with Todd that dissaeptus should be considered a synonym of palustris. As of the fifth edition of the A.O.U. Check-list (1957), therefore, the nomenclatorial status quo is as follows: Telmatodytes palustris palustris (Wilson), Long-billed Marsh Wren. Telmatodytes palustris iliacus Ridgway, Alberta Marsh Wren. Telmatodytes palustris dissaeptus (Bangs), Prairie Marsh Wren. Over two decades ago Welter (1935, p. 4) pointed out that the so-called Prairie Marsh Wren was actually a composite of two recognizable races, a western and an eastern, for which he used the names iliacus and dissaeptus respectively. His comments, which appear only in a brief foot-note in a general life-history paper, have been overlooked or ignored in subsequent years. Welter's diagnosis of the characters of the two alleged races appeared in more detail in his Ph.D. thesis (Welter, 1932), the taxonomic portion of which was never published. Only Oberholser (1938, p. 448), after Welter’s paper appeared, recognized both iliacus and dissaeptus as distinct subspecies, and he may have come to this conclusion independently, since he does not quote Welter. Oberholser used the vernacular name Prairie Marsh Wren for the western iliacus , and called the eastern dissaeptus Massachusetts Marsh Wren. There are three taxonomic points which obviously must be decided before the nomenclature of the northeastern Marsh Wrens can be straightened out. These questions may be listed as follows: 1. Was Todd correct in considering the type series of dissaeptus insepar- able from palustris ? 2. Was Aldrich correct in identifying the type specimen of iliacus as a migrant of laingi , the Alberta Marsh Wren? 3. Were Welter and Oberholser correct in believing that four rather than three races should be recognized in the area covered by this study? The treatment in the fifth edition of the Check-list, as given above, re- quires that questions 1 and 3 be answered “no”, and question 2 answered “yes”. Believing that, as with so many other seemingly complex taxonomic prob- lems, the solution lay in assembling, for once, a really adequate series of specimens, I was able to bring together approximately 670 skins of Tel- matodytes palustris representing only those northeastern populations under discussion here. Museums and individuals from whom specimens were bor- rowed are listed in the first paper of this series (Parkes, 1955). The three questions listed above will be taken up one by one, and then a summary of the characters and distribution of the subspecies I recognize will follow. 1959 Parkes: Systematic Notes on North American Birds 277 1. The status of the name dissaeptus Bangs Among the specimens examined during this study were two topotypes and a number of near-topotypes of dissaeptus. Comparison of this series with a near-topotypical series of palustris (vicinity of Philadelphia, Pa.) indicated that the Massachusetts birds were nearer the inland race as exemplified by specimens from central and western New York. There is some intermediacy toward palustris, but I feel that the relationships of the Massachusetts population to which Bangs applied the name dissaeptus are definitely with the inland form. Much of the earlier confusion, including Todd’s rejection of dissaeptus as a synonym of palustris, can be traced to the use of speci- mens from the vicinity of the District of Columbia as typical of palustris when making comparisons. Unfortunately there are far fewer specimens of good palustris in museums than there are of these District of Columbia birds. The latter show much intergradation with dissaeptus and also with waynei, a race of the coastal marshes of southeastern Virginia and North Carolina which is outside the scope of the present paper. When topotypes of both palustris and dissaeptus are compared, as mentioned above, it be- comes evident that the latter name is not a synonym of palustris, but is available for an inland race. 2. The status of the name iliacus Ridgway I have examined the type specimen of Telmatodytes palustris iliacus Ridgway (U.S.N.M. 90199, from Wheatland, Indiana). It is most unfor- tunate that this particular specimen should have been chosen as a type, as it may or may not have been a migrant when collected (April SO, 1883). After careful comparison with excellent series of both populations to which this type has been assigned (the Prairie and Alberta Marsh Wrens), I have come to the conclusion that this specimen does not represent a migrant of the Alberta race. It is admittedly paler than the great majority of Prairie Marsh Wrens, but it is by no means as pale as laingi, nor are the black por- tions of the plumage suffused with light brown as in typical laingi. The specimen represents either a light-colored extreme or (as Mr. Todd believed) a foxed specimen of the Prairie Marsh Wren. Geographic evidence is even more pertinent; I have seen no migrant specimens of the Alberta Marsh Wren taken east of Oklahoma, and it is apparently uncommon even that far east. The chances of the Indiana-taken type of iliacus being an actual derivative of the Alberta population are, from the geographic standpoint, infinitesimal. The Alberta Marsh Wren thus reverts to its proper name of Telmatodytes palustris laingi Harper, as listed in the fourth edition (1931) of the A.O.U. Check-list. With the earlier name dissaeptus available for the inland population that has been called Prairie Marsh Wren, the fate of the name iliacus depends on the answer to question 3. 3. The number of recognizable northeastern races Examination of the excellent series of specimens assembled for this study convinced me that Welter and Oberholser were perfectly correct in their assertion that two subspecies are involved within the range described for dissaeptus in the fifth edition (1957) of the A.O.U. Check-list, making a 278 Annals of Carnegie Museum vol. 35 total of four within the area encompassed by this study. Names are avail- able for all races (palustris, dissaeptus, iliacus , laingi), and no major change of concept of the names as used in the fourth edition (1931) of the A.O.U. Check-list is necessary other than the revival of iliacus Ridgway for the western component of the “Prairie Marsh Wren”. A summary of the characters and distribution of the four recognized races follows. Telmatodytes palustris palustris (Wilson) Certhia palustris Wilson, American Ornithology, 2, 1807, p. 58 (borders of the Schuylkill and Delaware rivers at Philadelphia, Pennsylvania). Characters: All browns, especially on sides of neck and flanks, more grayish or olivaceous, less reddish or buffy than in other races; browns of back and rump more variable, but show the same tendency. Crown the blackest of the four races, with the brown central stripe often reduced. When present, this crown stripe is almost always well defined, not diffuse. Lores white. Black and white interscapular region averaging somewhat more extensive, with broader white streaks. White superciliary stripe well marked, breaking up at posterior end into white dashes which tend to be continuous with white dashes of interscapular region. A brown collar seldom present between black of crown and of interscapular region. Throat and breast usually immaculate white. Slightly smaller and bill more slender than inland races (see measure- ments). Range : Strictly a coastal form; all inland specimens I have examined from eastern United States north of the Florida peninsula are referable to dis- saeptus or iliacus no matter how they were originally identified. This in- cludes, for instance, such specimens as that from Berea, Kentucky (which I have examined), identified as palustris by Wetmore (1940, p. 549). Nominate palustris breeds, in general, in coastal and estuarine marshes from Rhode Island to Virginia. Intergrades with dissaeptus in Connecticut (presumably), the Hudson valley, western New Jersey, and southeastern Pennsylvania. Birds from Chesapeake Bay average nearest palustris; those from Washing- ton, D.C., and adjacent parts of Virginia and Maryland are variously inter- mediate between palustris and dissaeptus , with some individiuals showing an approach to waynei, the next race to the south along the coast. Inter- gradation with waynei manifests itself most conspicuously in the presence of barred upper tail coverts. Winters in much of its breeding range, and south through most of peninsular Florida. Accidental in Nova Scotia. Telmatodytes palustris dissaeptus (Bangs) Cistothorus ( Telmatodytes ) palustris dissaeptus Bangs, Auk, 19, 1902, p. 352 (Wayland, Massachusetts). Characters : All browns more reddish than those of palustris. Brown always present on crown, almost always diffuse, often covering almost all of crown, frequently extending backwards and laterally as a collar between black of crown and of interscapular region. Lores usually white, sometimes pale buffy. Black interscapular region averaging somewhat less extensive, white markings narrower than in palustris. Superciliary line often less pure white, less distinct. Postocular spot or line usually brown (black in palustris). Brown of sides averaging more extensive, especially anteriorly. Breast frequently washed with pale brown or orange brown. 1959 Parkes: Systematic Notes on North American Birds 279 Range : Breeds from southern New Brunswick, southern Maine, Vermont, New Hampshire, southwestern Quebec and southern Ontario south to the mountains of western Virginia, eastern West Virginia, south-central Penn- sylvania, and Ohio. Iritergrades with palustris just inland from the coast from Rhode Island to Virginia, and with iliacus in central Michigan; un- doubtedly also intergrades with iliacus in western Ohio, eastern Indiana, and south-central Ontario. Winters at least casually through most of the breeding range, south to the Gulf states from Mississippi eastward. Accidental in Nova Scotia. Telmatodytes palustris iliacus Ridgway Telmatodytes palustris iliacus Ridgway, Proceedings of the Biological Society of Washington, v. 16, 1903, p. 110 (Wheatland, Knox Co., Indiana). Characters'. All browns lighter and more reddish than in dissaeptus. Crown almost always more suffused with brown than in that race. Lores and an- terior portion of superciliary line usually buffy, sometimes white. Ear covert region more rusty. Brown of crown tends to be more extensive posteriorly. Light markings of outer webs of secondaries and tertials more extensive, tending to be confluent in some specimens. Always at least slightly stained with rusty below, often most of the underparts so washed. Range : Breeds from southwestern Michigan, western Indiana and western Ontario south to southern Illinois, Missouri and eastern Kansas, west through the Dakotas and Manitoba. Most Manitoba specimens are of this race; some from the western part of the province show the influence of laingi. As is also the case with dissaeptus , the type locality of iliacus is near the eastern periphery of the range of the subspecies, rather than near the center of differentiation. The characters of the race to which the name iliacus is being applied are shown to best advantage in specimens from the upper Mississippi drainage, particularly Minnesota and Wisconsin. This subspecies winters through much of southeastern United States, west at least to southeastern Texas. Like many other midwestern birds, iliacus has a southeast-northwest migration route, and may be found on migration within the breeding range of dissaeptus , and casually as far east as Long Island, New York, in the range of palustris (A.M.N.H. 66451, Montauk Point lighthouse, October 13, 1888). Telmatodytes palustris laingi Harper Telmatodytes palustris laingi Harper, Occasional Papers of the Boston So- ciety of Natural History, v. 5, 1926, p. 221 (Athabaska Delta, Main Branch [9 miles above mouth], Alberta, Canada). Characters : A very distinctive race, with all browns paler and buffier than in iliacus. All black areas of other races, including interscapular region, are in laingi more or less invaded by browns. Under tail coverts immaculate or barred with pale rusty, never with black barring ( iliacus almost always has at least a trace of black in the barring, which is often heavy). Differs from plesius and other western races in lacking bars on the upper tail coverts. Within the range here assigned to laingi , there is some geographic variation in size. Topotypes average slightly longer in wing and tail than any of the other races discussed here. Saskatchewan birds average somewhat smaller in these measurements. In the accompanying table, after the column giving 280 Annals of Carnegie Museum vol. 35 measurements of a composite series of male laingi comparative wing and tail measurements are given for a small series each of male topotypes from Alberta and males from Stalwart and Last Mountain Lake, Saskatchewan. The only female laingi measured were from the latter series. Range: Breeds east of the Rocky Mountains in Alberta. The breeding Marsh Wren of Montana east of the Rockies is probably this race (no specimens seen), and is so assigned by the A.O.U. Check-list (1957). Birds of all but easternmost Saskatchewan are closest to laingi in color, but may show some signs of approach to iliacus. Presumably intergrades with plesius along the east face of the Rocky Mountains; no breeding birds from this area seen, but winter and migrant specimens which appear to be laingi-plesius inter- grades have been examined from New Mexico, Texas, and Mexico. Prin- cipal winter range apparently in Mexico; migrants examined from Colorado, Oklahoma, New Mexico, and Texas. REFERENCES Aldrich, John W. 1946. New subspecies of birds from western North America, Pro- ceedings of the Biological Society of Washington, v. 59, p. 129-136. American Ornithologists’ Union 1931. Check-list of North American birds, ed. 4. 526 p. 1957. The same , ed. 5. 691 p. Oberholser, Harry C. 1938. The bird life of Louisiana. Louisiana State Department of Conservation Bulletin 28. 834 p. Parkes, Kenneth C. 1955. Systematic notes on North American birds. 1. The herons and ibises (Ciconiiformes). Annals of the Carnegie Museum, v. 33, p. 287-293. Todd, W. E. Clyde 1937. Critical remarks on the Long-billed Marsh Wren. Proceedings of the Biological Society of Washington, v. 50, p. 23-24. Welter, Wilfred A. 1932. The natural history and taxonomy of the Marsh Wren Tel- matodytes palustris (Wilson). Cornell University Ph.D. thesis, unpublished. 167 p. 1935. The natural history of the Long-billed Marsh Wren. Wilson Bulletin, v. 47, p. 3-34. Wetmore, Alexander 1940. Notes on the birds of Kentucky. Proceedings of the United States National Museum, v. 88, p. 529-574. Wetmore, Alexander (Chairman) and others 1945. Twentieth supplement to the American Ornithologists' Union Check-list of North American birds. Auk, v. 61, p. 436-449. 1948. Twenty-third supplement to the American Ornithologists’ Union Check-list of North American birds. Auk, v. 65, p. 438-443. TABLE 1. MEASUREMENTS OF TELMATODYTES PALUSTRIS 1959 Parkes: Systematic Notes on North American Birds 281 j© Ch' .5 eg S 3 5 q J XO — 'CM > — 00 xO M 5 ^rr- *^3 © nt< © -© PS V3 i— H CO I>* ■Ofi oo CO °° ^oo CO 00 •CO GO • v^wi ■ w cm d *> oo £ 00 ' ' iO O CM CM CM^o CM CO © •«> pj °9 (S' ^ ^ fX 10 ^ ^ ■S < >9 S ^ !3 © ^ CM xO 'Sb 8 g- g s- •5 S«S«$2eSS3« >— i -W-CM '«-'© v — . © xO ^ ' © CM 00 '©■GO a ” © cm ph ^ cm .53 H^^trJH^HCOooX^ ^2 xo '^'co -w© - © ^ xO © © CO © © CM ^9 • oo CM «? s T i— h vi^CM •©'CD ’’f xO 2© © S «9 © & © CO o 00 *5 CM -©*© ©“• CM '©'t" '©'CM f©' 2 °° -- • 0O • xfj CQ 50 ^ nfs ^ ^ © i >-h co go c5 iu 0O oo •^'I>» '-'XT5 n_- I © £ S o t3 ^ co « © xo ^ijo ;©'xr> j©eo ^©'© CO wi4» s-'l> © oo CM 00 S”® ,«2 2 oo ^ <* « 3^2© ^ c* ?? cm © '© ^XJO ®° © CM CM ,-•00 -©- I— I ^ <® CM ^ '"T »? ^ ^ ^ i— t GO ' — ^© xfj © GO -©-xo ©no © • © CM © CO CO cO o ^ "-"© 00 CM | , 1/3 PS ■5 CO U *5h •*-» CO O p0 x-> PS bo fl