Hike with a Dendroecologist Part 2

The road narrowed and grew shrubby as Andrew, Maddie (a graduate student at OSU) and I pulled up, looking for a place to park. Just behind us was the rest of the team in a second vehicle—Charles, another experienced member of the crew, at the helm. 

“This one is very steep,” said Andrew with a grin. “We are going to challenge some of them [crew members].”

We were at Site number 93. The area had been sampled for fire scars on the opposite side of the road already. The plan today was to focus on setting up a plot and gathering forest development data which mainly involves coring trees. 

Forest development, Andrew explains, is the history of how a forest changes over time. By gathering plot data and coring trees, he hopes to understand the different development trajectories a forest might undergo in relation to its disturbance history. 

Fire Record

The fire record for the site showed fires in 1759, 1836, 1844, and 1883. But you wouldn’t know it from looking at it. The trees at this site were much larger and a good deal older than the forest we had sampled earlier in the day.

The fires, Andrew predicted, must have been low severity with low tree mortality since there are so many older living trees at the site. The oldest of the trees are over 800 years old, established in 1190, possibly following a stand-replacing fire.

“Now when we core the site,” said Andrew, “It will be interesting to see if we find cohorts of hemlock and cedar related to these fires [in the 1800s] because presumably they come in because of gaps or openings.”

Plotting a Plot

After gathering all of the necessary equipment, the crew, Andrew and I headed down through a thicket of young conifers that lined the road. The slope was steep as we navigated our way over and around downed logs and branches. Eventually, we reach the ancient forest and what would be our plot center.

Andrew’s plot design follows the same format as the Forest Inventory Analysis Program of the U.S. Forest Service—with three circular plots that branch out at 120 angles from a central plot (each with a radius of 58.9 feet) and smaller subplots and microplots inside these.

Andrew staked out the center of plot one. He chose the site in hopes of capturing multiple cohorts of trees, including a large old-growth Douglas-fir that sat just downhill of the center.

While he gathered some basic plot data, he set most of his crew to work on coring trees.

Growth Rings

Gathering core samples from all the trees in the plot is both the most time-consuming and important part of the plot data Andrew’s crew collects. Each core is carefully packaged and carried back to the lab for closer analysis.

Cores can tell you a lot of things about trees. The size of the rings tells you about growing conditions, while the number of rings tells you about the age of each tree and when it was established. 

Charles and some of the crew stayed nearby coring some of the smaller trees on the uphill slope. Maddy and one other crew member took a stab at the behemoth below.

Coring Efforts

I sat with Charles for a while to try and learn the secret to getting the perfect core. Charles was coring a hemlock tree—a species notorious for being rotten on the inside. The bit screeched as he rotated the handle.

“As a general rule, if a tree is leaning the pith should be away from the lean,” Charles explained. “I chose the side that is kind of oval. I want to core into the shortest side of it… to core the shortest distance.”

Charles originally studied the Classic Period in school before returning to graduate school to study forests—something he had wanted to do for years. He explained enthusiastically how the field of forest ecology has placed a greater emphasis on humans as part of the ecosystems in recent years. A change for the better.  The borer clicked as he talked.

Before long, Charles reached the center and stopped coring. He rotated the handle counterclockwise, breaking the connection between the core and tree, and gently removed it on a thin metal spoon. Unfortunately, there were tiny holes in the wood indicating rot.

“I might try a few more turns,” he said.

Persistence

As Charles continued coring his hemlock, I dropped down to check out what progress was being made on the old-growth Douglas-fir at the bottom of the hill. There was likely 800 years or so of rings to get through. I couldn’t imagine the effort it would take.

The team had just extracted a couple of massive pieces of core from the tree and laid them down on a nearby log. They looked pretty good to me, but Maddy wasn’t satisfied. The core was still short of the pith.

They tried a new angle and took turns spinning the bright blue handle. Gloves were a must-have for this line of work. And work it was. Maddy and her partner would spend the remaining part of the day attempting to get a quality sample from the tree.

Old Growth

With the creaking and squeaking of borers in the background, I decided to take a little hike around the forest plot. I wanted to get a different perspective of the forest and see if I could get a view of some of the old growth treetops.

When we first arrived at the site earlier in the day, Andrew was quick to point out that despite their size, most of the trees downhill from us were not old growth. Most of the trees we were able to see from the road still had a tapered top, having not reached full height or had time to flatten out.

In fact, most of the trees were probably about 260 years old—middle-aged for a tree—establishing after the 1759 fire.

“A 500-year-old tree will be really wide at the top,” Andrew explained, “and will have wide branches…”

In short, a flat-topped tree is an old growth tree.

Balancing near a log, I took a picture of a couple gnarly looking old growth trees and their heavily branched tops.

Old Assumptions

Work continued under the tall canopy. There were a lot more trees to core and measure.

As the crew worked, Andrew told me that the data we were collecting that day might also be used to test an old assumption about old growth trees. 

“One of the assumptions about old growth forest is that it has no net change in biomass,” Andrew explained. “Whatever is dying is being balanced by what is growing.”

Old growth is sometimes romanticized as a stable, unchanging system. This may be true, but as far as Andrew is concerned there is a lack of evidence to say one way or the other. Based on careful reconstructions of the history of old-forests it seems more likely that they are always changing. Sometimes change is quick, like after fires or other disturbances, and at other times slow, at an imperceptible pace for our relatively short lifetime.

“Coring all these trees, we can quantify basal area increment over time,” said Andrew. Basically, you take the width of the rings and diameter of the tree to determine the volume of wood added each year. 

This sort of data could answer a myriad of questions:

Does biomass increase or decrease in these old growth stands? Or, is the answer, it depends? How quickly are the younger cohorts replacing older slower growing trees and trees that have died? Where is the stand headed in terms of what types of trees and their structure?

Only time and an increment borer will tell.

Other Measures

Of course, there is a lot more data to collect than just taking cores. For each tree in the plot, DBH (diameter at breast height), height, species, and condition are recorded. Then, there is also a host of other measurements, like slope and aspect of the plot.

As I returned from my romp through the forest, I caught Andrew taking another measurement—woody debris.

Curious, I asked what he hoped to learn from tracking the downed wood in the plot.

First, he explained that it can relate to disturbance history as you compare what is on the ground with what material is consumed.

“Different development histories produce different woody debris,” he suggested.

“Honestly,” he added, “I don’t know how it will turn out.” However, if some sort of pattern does emerge in the data, by using the same protocols as FIA, it opens up possibilities for access to an even larger data set.

Back Again Tomorrow

It was late afternoon when Andrew’s crew wrapped up for the day. They probably would have stayed out even longer if I didn’t need to get home.

On the drive back to my car we chatted about all sorts of things—family members, hiking and climbing, and of course research.

Andrew and his team are not done. They will be back tomorrow. And probably the day after.

How does a forest grow?

In a million different ways.

Ask Andrew and his crew—they have all sorts of theories to test.

Andrew Merschel is an ORISE postdoctoral fellow working with the USFS PNW Research Station and he leads the tree ring lab at Oregon State University. Andrew uses tree rings to develop a shared understanding of how different forest ecosystems function over time. He is particularly interested in how disturbances (mostly fire) and forest management have shaped and will continue to shape forest ecosystems in the Pacific Northwest.  Andrew lives with his family (Vanessa, Aldo, and Sawyer) in Corvallis, Oregon and they enjoy a mixture of fishing, hiking, wildlife ecology, and chainsaw repair in their spare time.

Forest Hike with Bird and Wildlife Biologist

Rushing water. A shushing breeze. Rustling leaves. Chattering wildlife. These are the sounds of a forest in the foothills of the Willamette Valley. Soft, tranquil, quiet. Or at least in winter.

The forest awakens in spring. As flowers stretch out their petals and leaves unfurl to catch the sunlight, the tranquil chatter of the forest turns into an all-out symphony of sounds. Like the string section in the orchestra, it is the birds that draw the most attention.

I have always enjoyed bird song but have not yet mastered their melodious rhythms. This spring I am determined to take a closer listen.

Fortunately, Joan Hagar, a research wildlife biologist with USGS, agreed to meet with me to talk birds in a local forest.  

The Hike

  • Trailhead: 720 Gate at the end of Sulpher Springs Road
  • Distance: approximately 2 miles
  • Details: Limited parking at the end of a well-maintained gravel road. No fee for parking. No restrooms. Park at gate 720 gate and head up Road 720. Look for a right turn-off on a user trail that takes you back to the gate. Map of area available on OSU College of Forestry website.

Introductions

I met Joan on a cool spring afternoon. It was overcast, but not raining. Would the birds be out?

We didn’t take but a moment before heading up the trail which rose along a riparian corridor next to a rushing creek.

I asked Joan to tell me more about herself and her career.

“The focus of my career has been to help forest managers incorporate wildlife habitat into their management plans,” she explained as we walked. “Remind them that they can accommodate wildlife at the same time as they are meeting their other goals.”

More specifically, she is all about the birds. Joan has spent her career studying birds and other wildlife in the Pacific Northwest.

As Joan explained it, she was born with it.

“My dad was a wildlife biologist and taught me the birds,” she explained, “and being able to hear them and know what species you are hearing it is like understanding a foreign language.”

A skill she would prove multiple times on our walk, but at least for the moment, the forest was rather quiet.

Indicators

As we continued our gradual climb up the forested hillside, I asked Joan “Why birds?”

“Birds, it turns out, are really great indicators for management and environmental change,” explained Joan.

Many species are only suited for a particular habitat or forest type. If the environment changes, so does the bird community. As a master’s student, Joan explained, she was able to see this firsthand. 

Joan studied the impact of forest thinning on bird communities.

“I am going to show that harvesting is bad for wildlife,” Joan’s early scientist idealistic self-had thought, but she was mistaken.

“I found out that when the canopy of these dense conifer stands opened up and allowed the understory to develop… that meant more productivity—more flowers, fruits, seeds, and insects,” said Joan. 

In essence, thinning increases resources birds relied on and as a result bird diversity also increased as birds that were attracted to the more open habitat arrived.

“Disturbances aren’t a bad thing,” Joan concluded. 

Of course, “that is a bird perspective,” said Joan. “Amphibians might feel differently.” 

Why birds?

In addition to birds’ ability to respond so quickly and clearly to environmental change, there are many other reasons birds are useful biological indicators. 

“Birds are everywhere,” said Joan. “And they are fun to watch.”

Joan tried studying amphibians early in her career but found it more difficult.

“You have to turn over a lot of logs to find them,” Joan explained, “and in doing so you have to destroy their habitat.” 

(Turns out, Kermit is right—It ain’t easy being green.)

Birds, on the other hand, can be counted by sight and/or sound.

For more detailed demographic data, mist nests may be used to capture the birds temporarily to study them. By using a method called “mark-recapture,” even the abundance of birds may be calculated.

Riparian Resident Birds

Deciduous trees, like bigleaf maple and red alder, having still not leafed out, offered views down towards the water as we walked. 

“So, what kinds of birds would you find here?” I asked.

“Usually there are a lot of birds here,” Joan responded and pointed out the chattering call of the Pacific Wren.

“They [Pacific Wrens] start nesting this time of year,” she continued; “they like a lot of dead wood—stumps, logs—and they love the riparian area because of all the trees that fall in and it is damp and moist.”

Pacific wren is a resident species in Oregon’s western forests, along with Spotted Towhee, Song Sparrows, Canada Jays, and Steller’s Jay.

Barred owls and Pygmy owls are also common residents found nesting in snags.

“I have long suspected a Pygmy Owl nesting near here,” said Joan.

Riparian Breeding Birds

“In a normal year we would be hearing warblers,” Joan continued as we rose above the creek.

Orange-crowned Warblers usually arrive in April, with Hermit Warblers arriving a few weeks later.

“They [Hermit Warblers] are really cool because they only breed along the west coast here—from the coast to the Cascade Mountains,” said Joan excitedly.

Hermit warblers are what Joan called “endemic breeders.” Traveling to Central America during the non-breeding period and returning to their narrow breeding range in Pacific Northwest forests.

“Pacific-slope Flycatcher,” Joan recalled is another riparian migrant. “I am usually starting to hear those this time of year.”

Pacific-slope Flycatchers are especially fond of forests and woodlands near waterways where the canopy is dominated by deciduous foliage—often nesting on the slopes of forested canyons.

“They love these riparian trees, like maples and ash,” Joan remarked. Here the flycatchers catch insects below the canopy.

Woodpeckers

Early spring is also a great time to see woodpeckers in Oregon’s Willamette Valley forests.

“Hairy woodpecker, Downy woodpecker, red-bellied sapsucker…” Joan rattled off some examples.

It is nesting season and woodpeckers are out scouring the woods for the perfect tree to build a nest in.

“Woodpeckers are primary cavity nesters,” Joan accounted.

Primary means that they excavate their own cavity, as opposed to secondary cavity-nesters, like chickadees, bluebirds, and wrens, that depend on woodpeckers to provide cavities.

“They do the excavation of the cavities because they have strong bills,” Joan explained.

“Woodpeckers are funny because they do a lot of excavating before they settle,” she continued. “The male goes around and makes a cavity, then the female checks it out and goes ‘eh’ and so he makes another cavity.”

This process continues for a while until the female is satisfied. Fortunately, the result is several new unoccupied cavities produced each nesting season. This is great news for secondary cavity nesters, like chickadees and nuthatches, who are soft-billed and reliant on finding a home in already existing cavities.  

“They [woodpeckers] are considered ecosystem engineers because they make habitat for so many other species,” explained Joan.

“So, if I see some sort of hole, it is likely something lives in there?” I asked.

“It’s likely,” Joan responded.

Preferences

Eventually, the trail bent and moved away from the creek, heading out on a slowly rising wooded ridge dominated by Douglas-fir.

Standing out in the mix of trees was the statuesque Pacific madrone, with its red shredded bark and green leathery broadleaves leaning out along the trail’s edge.

“In the fall, the madrones have a lot of berries and the band-tailed pigeons were feasting,” Joan reminisced. “They were covering the trees!”

Joan also noted how madrones tend to have cavities in live trees, unlike conifers that need to be dead or dying.

I asked Joan if certain species prefer certain trees.

In general, primary cavity nesters prefer hard snags. However, there also seem to be some preferences in terms of tree species.

“Pileated Woodpeckers like grand fir,” Joan offered as an example, speculating that perhaps it had to do with the decay process. And “Red-breasted Sapsuckers like maple trees,” frequently excavating a nest in a dead branch of a live maple.

Apparently, there is an entire branch of ecology that studies the relationship between primary and secondary cavity nesters and the trees they occupy. Joan mentioned “cavity-nest webs” as a way researchers aim to delineate and describe the complexity of these relationships.

In any event, there is one consistency—“good snags are scarce” and hard to come by.

Harvest Unit

Speaking of good snags, soon Joan and I crested the hill, we broke out of the forest into a clear-cut harvest unit littered with snags and potential snags.

“It is really nice to have something out here,” said Joan referring to all the trees that were left behind.

Joan has consulted on previous harvest projects and recommended that forest managers leave more snags and live trees than might be typical in a clear-cut.

Joan pointed to a large snag with twisted branches that had been left behind.

“That snag they left isn’t worth anything because it is gnarly,” said Joan referring to the potential timber value, “but for wildlife, it is worth a lot.”

Disturbance

Joan was also quick to point out that the clear-cut itself offered some benefits to wildlife.

“There are actually a lot of species that evolved with disturbance,” Joan remarked. “Disturbance is not a bad thing.”

Species like swallows, wrens, pigeons, Purple Martin, and a whole host of raptors benefit from the opening in the canopy.

“This is a phase of forest succession—early seral,” she continued. “When it is natural it is a very diverse stage.”

Unfortunately, it wasn’t all good news in the clear-cut, as many of the shrubs that come up during the early seral stage were sprayed with herbicide to give the next generation of conifers a competitive edge.

I was also struck by the small size of the clear-cut and asked Joan about it.

“Is it good to have smaller clear-cuts?”

“There is no one good size,” said Joan.

She explained that for a forest species having a small clear-cut makes the forests more permeable—a species that wants cover can go between trees. However, the larger the clear-cut, the more valuable the area is for a species that needs open areas.

“There is always a trade-off,” said Joan. Her advice for land managers—“be as variable as possible, and work with what is there.”

Ghost Forest

As we walked past the clear-cut with the intact forest on our right, it was easy to assume that the intact forest was in some way “natural” or “right.” But, as Joan reminded me, the conifer forest only exists on this hillside as a product of colonialism.

“Before the European settlers came,” explained Joan. “Native Americans burned this area—it was a bald with scattered oak and scattered Douglas-fir. It was very open.”

With colonialism came fire suppression and the conversion of oak woodlands and prairies into forests.

“If you look in this forest now, you can find old oak trees,” said Joan. “You can tell they are open grow with lateral limbs, but they are dead and decaying…”—overshadowed by Douglas-fir.

We looked deep into the thicket of forest for one of these “ghost oaks,” and found what looked like a mossy, dead limped giant of an oak tree.

“There used to be a bird species that used those,” remarked Joan. “Lewis’s woodpecker—iridescent green with a red breast—they valued the oak and ponderosa pine.”

She sighed, “Now, they don’t nest here. There is not the habitat for them.”

Purple Martin

Then we passed it—a white sci-fi-looking apparatus on the hillside to the left.

“Here is my Purple Martin gourd rack,” laughed Joan. “It is ugly as sin!”

However, what it lacks in aesthetics, it makes up for in function.

Joan explained that the rack is put up to provide a temporary nesting opportunity for Purple Martin—a threatened species here in the west. As insectivores, Purple Martin hunt insects on the wing, so in addition to needing natural cavities for nesting, they also need open space for hunting—a difficult combination to achieve these days.

“The public land has all the big snags but is too dense, and the private land has open areas but not the snags,” explained Joan.

The rack is meant to provide temporary housing until the woodpeckers can create the cavities in snags Purple Martin needs.

However, she cautions people from putting up their own gourd racks. The eastern population of Purple Martin are entirely dependent on people for nesting for this reason. She wants to avoid this in the West.

“Purple martins are the poster child for snags,” she proclaimed.

 Across the clearing, I saw a small cavity in a Pacific Madrone. I asked Joan if that might work for the Purple Martin or some other species.

“It looks good for a pygmy owl,” she replied, “but I am not sure they would want to be out in the open. A flicker would love it,” she laughed. 

What about Yew?

We were nearing our turn off into the woods when we happened past a shaggy-looking Pacific Yew.

“They always make me think of old forests,” Joan smiled.

“Does it do anything for wildlife?” I asked.

“I don’t know anything in particular,” Joan replied. “They are good for cover,” she offered.

What about Joan? We knew what the Yew was up to (being a really cool tree!), but what about Yew? I questioned Joan, pun intended.

 “Right now, I am working on Purple Martin stuff,” she said—tracking them with GPS in collaboration with Klamath Bird Observatory and trying to figure out where they go in winter. So far, she has found that they spend some time in Baja—sounds pretty good to me.

“That is one thing,” she said. “I am trying to finish a bunch of projects,” Joan confessed in preparation for retirement before the end of the year—that also sounds pretty good to me. Maybe she will have to visit Baja?

“Another project is not birds,” she continued, but a carnivore survey using camera traps in the Klamath Network of National Parks.

“We are looking for Marten, Fisher, and Sierra Nevada Red Fox,” said Joan.

She explained that there is a lot of interest in carnivores. They are not only sensitive to environmental change and have been facing declining population rates, but they are also an important part of the food web.

Dense Woods

We were on the steep downhill return trail when I spotted a large patch of Oregon Grape out of the corner of my eye. 

“Do they help birds?” I wondered out loud.

“I don’t know,” Joan responded thoughtfully. “The hummingbirds love the flowers.”

Soon we were considering the Oregon Grape fruits and species that might benefit from them as a food source as well.

In the distance, Joan heard the call of a Kinglet deep in the woods. Kinglets, she told me, were birds that responded negatively to thinning in her graduate research.

“They are beautiful little birds,” she described. “A bright gold crest with a scarlet, orange stripe down the middle.”

She heard the call again—“high and thin.” Whatever she was hearing, I didn’t register.

Learning Birds

“Is it hard to tell birds apart?” I asked.

“Not for me,” she laughed. “But yes.”

So how does one learn? Joan had a few tips.

First, “Come during the off-season,” she suggested. Learn the birds that are common year-round and learn them one at a time.

Second, she recommended using an app, like the Merlin App to help, as it identified with sound, and you can get the results often right away.

Finally, get a feeder. Feeders are an excellent way to meet several of the birds that are around all the time.

Some starter birds include song sparrows, dark-eyed junco, chickadees, nuthatches, and towhees.

It also doesn’t hurt to have a bird with a favorite song. Sometimes that is enough to draw one in. 

“My favorite is the hermit thrush,” said Joan—a high-elevation bird with a song. “It sounds flute-like and ethereal.”

I recalled hearing the bird myself while hiking in the Jefferson Wilderness—singing its heart out well into the evening. Afterward, I had to find out what I was hearing!

Help the Birds

The trail continued down through the dense forest before dropping us back on the wide gravel road we had come up on—back in the riparian forest.

As we made our way back down to our cars, I asked Joan if she had any tips for helping birds.

“Audubon has a list of 10 things you can do for birds,” Joan responded.

“The biggest problems are hitting windows, lights during migration, and cats,” she continued.

So, to help with that, she suggests putting bird strike prevention on any windows that might fool birds, turning out the lights during migration, and keeping pet cats indoors.

Now, with advancements in bird tracking, you can find out when birds migrate through your area, so you know when dark skies are most important.

Pesticides are another concern she brought up.

“Anything that affects insects affects birds.”

Brown Creeper

“Well, we didn’t see very many birds,” Joan remarked when were just about at our cars.

Then, she spotted something up in the trees—a small brown bird hopping up the trunk. It was a Brown Creeper.

“They go way up and then they fly down to the base of the tree or their nest,” Joan noted. 

I watched the Brown Creeper hop its way up a large Douglas-fir trunk before taking flight and landing on another tree nearby.

It was probably feeding on spiders hidden in the bark or collecting web for its nest—a common practice according to Joan.

The light was dimming as we stood and looked up at this small brown bird doing what it does best before we lost track of it.

Trills and Thrills

“That was fun!” proclaimed Joan.

And I too felt satisfied.

We have only heard or seen a few birds, but I was walking away with more bird knowledge than I could have imagined.

High-pitched trills spilled through the trees, like a tumbling stream, as we walked the last few feet to our cars.

And I knew it was the Pacific Wren singing us off.


Joan Hagar is a Research Wildlife Biologist with the U.S. Geological Survey. She has been studying birds and other wildlife professionally for the last 30 years.

Hike with a Geologist at Barnes Butte

View up to Barnes Butte from the trail

At first glance, a visit to Barnes Butte in Prineville looks a lot like much of central Oregon—a landscape of sage brush, juniper, and volcanic rimrock. It is difficult to imagine that Barnes Butte is, in fact, the inside edge of a massive supervolcano that—though now extinct—erupted more than 240 cubic miles of material forming a caldera roughly 29.5 million years ago.

Approximately 25 miles by 17 miles in size, the oblong-shaped Crooked River Caldera reaches from Smith Rock State Park in Terrebonne east to the Ochoco Reservoir and south to the Prineville Reservoir and Powell Buttes. For something so large, it might seem surprising that it wasn’t until 2005 that a couple of scientists first noticed its presence.

However, standing in the parking lot of Barnes Butte City Park with Carrie Gordon, a retired geologist, and willing hiking partner, it became obvious why such a large geological structure went unnoticed for so long. Seriously, what volcano?

The Hike

  • Trailhead: Barnes Butte Trailhead
  • Distance: Varies. (2.7 miles w/565 feet elevation to top)
  • Details: Large parking area; no pass required; No restrooms (port-a-potty may be available)  

Introductions

It was a warm early fall day when I met Carrie in the parking lot of Barnes Butte City Park. Wildfire smoke created a haze across the skyline, but you could still just make out most of its features, including the Cascade Volcanoes in the distance.

Carrie, a small energetic woman, was all smiles as we gathered at her vehicle for introductions.

“I worked 40 years for the Forest Service,” Carrie said, “As a forest geologist.”

She explained that her job mainly entailed keeping track of material sources, like gravel.

“It is one of those careers that are just a hoot and a half,” she exclaimed.

Yes, this is Carrie. And we were just getting started.

Tall Tales

I asked Carrie to tell me about where we were standing.  After all, I couldn’t see any so-called “volcano.” She quickly pulled out her geology maps from her vehicle to orient me to the space and began to weave the tale.

“Jason McClaughry and Mark Ferns from DOGAMI started mapping in 2005,” she said. Originally, “they were supposed to map a 7.5-minute quadrangle,” Carrie continued.

Plans quickly changed, however. McClaughry and Ferns were tasked with finding water resources for Prineville, but while mapping, certain geological features started reshaping their goals.  By the end of the project, they had mapped over 903m2—and reshaped our understanding of central Oregon geology.   

“The cool thing about geology,” Carrie began, “The rocks don’t change but the story changes. We add to our body of knowledge, and we can go, ‘oh okay’…”

Anatomy of a Calderas

Perhaps the most important change to the story that McClaughry and Ferns brought to light was the chapter on the Crooked River Caldera. 

“Calderas are a little sneaky,” said Carrie.

Unlike, the very conspicuous Cascade peaks, “seeing” a caldera requires reading the landscape very differently. They are not peaks, rather, Calderas are mostly depressions.

Carrie explained: “Basic caldera formation is you have magma that is coming up to the Earth’s surface to the point you get a collapse.”

In the case of the Crooked River Caldera, these eruptions took place from about 29.7 to 27.5 million years ago. These were massive eruptions of rhyolitic lava, including volcanic tuff, that created a void below the volcano that eventually collapsed creating a 26 by 17-mile depression.

In addition, a ring fracture develops during caldera formation—allowing rhyolitic lava to intrude and bulge up along the side of the collapse.

Evidence of the ring fracture of the Crooked River Caldera can be seen at places like the Prineville Reservoir and Peter Ogden Wayside, where older rock that pre-dates the eruptions is tipping toward the interior of the caldera.

In addition, and perhaps even more obvious, rhyolitic domes can be observed marking the Crooked River Caldera Boundary. Carrie pointed to each—Powell Butte, Gray Butte, Grizzly Mountain, and, of course, Barnes Butte.

“This is the evidence that they [McClaughry and Ferns] found,” Carrie stated.

Looking out toward Gray Butte and Grizzly Peak (photo credit: Carrie Gordon)

Tuffs

I was beginning to see it—with many of the peaks visible from the parking lot—the caldera was taking form when Carrie whipped out another visual aid.

“I brought my box of rocks too,” she proclaimed.

Carrie pulled out two rocks with large flecks of material embedded within them—tuffs, I would soon find out.

“The cool thing about tuffs is they tell you about volcanic activity,” said Carrie. Tuffs are commonly associated with large violent eruptions as you see in caldera-forming.

“Tuffs are formed from bits and pieces of pumice and bits of rocks as it comes up through, in our case accreted terranes,” during an eruption, said Carrie.  “It is a mishmash of stuff.”

Pulverized stuff mostly, like ash, but also some solid flecks of rock, like pale gray pumice, embedded in the matrix—that is tuff. 

“It sparkles at you due to the crystal fraction in the ash,” described Carrie holding up two samples, her eyes sparkling more than the rocks.

Tuffs are also lighter than other forms of igneous rock, like other forms of rhyolite and basalt, as they are full of air pockets. She handed me one of the tuffs to weigh in my hand and basalt in the other—yep, I could feel the difference.

If you ever visited Smith Rock State Park, you have seen tuff. It is the tuffs that people mostly climb on. 

“Easy to pound in your pins,” Carrie remarked.

Tuffs from the Crooked River Caldera

Geochemistry and Cooling

Carrie had other rock samples in her box. She pulled out a shiny, black rock called obsidian, and a striped rock called banded rhyolite.

“These are all rhyolite geochemistry,” said Carrie. “Rhyolite has higher silica content than basalt and it tends to be blocky when it chills.”

However, the similarities end there.

“The thing about rhyolite is it comes in so many different forms.”

Tuff is the result of violent eruptions that pulverize rock, while obsidian and banded rhyolite are both formed as lava flows.

Obsidian is glassy because it cooled quickly enough that crystals were unable to form. Banded rhyolite, on the other hand, forms crystals that capture the layering that often occurs as lava flows.

“This is what makes up Grizzly and Gray Butte…” Carrie added, holding up the banded rhyolite.

She continued, holding up the two tuffs she had pulled out originally.

Tuffs to the left and obsidian to the right

“These are the same rock,” she explained. Only one had undergone a form of hydrothermal alteration, turning it “pistachio green,” while the other more “beigy” rock had not.

“And that is tuff,” Carrie concluded, putting her rocks back in her box.

She also mentioned granite—another form of rhyolite formed by a slow cooling process under the Earth’s surface.

“It is the same composition as obsidian,” Carrie reiterated, but “buried a long time.”

Just one more reminder to not take your rhyolite for “granite” (pun intended).

More samples from Carrie’s box of rocks

Off to the Races

At this point, we had been chatting for about 20 minutes and decided it was about time to hit the trail. The trail system at Barnes Butte City Park is rather extensive, but we kept it simple and headed up the Jockey Trail that goes along the base of Barnes Butte—an old trail that the landowners used to run horses on. 

As we started off on the rocky, dusty path, Carrie told me about the other trails that run through the park.

Apparently, much of the land was an old ranch. In addition to hiking the old horse track, there are also a lot of old cattle trails that are now hiking/biking trails that run through old grazing fields and around what used to be an irrigation pond.

Before that, there was even mercury mined on the Butte for a short time.

“See the main draw,” she said looking up toward the butte, “ there is an old BLM road that goes up to where the mercury mines in the 1940s are…. [The mercury mine is] courtesy of the caldera and volcanism.”

Mercury, lead, and gold, as well as Oregon’s state rock, the thunder eggs, rely on silica-rich waters to concentrate and form these minerals.

“You can take a footpath to the top of the butte,” Carrie added, “there are a lot of options.”

Rivers in the Sky

Soon we arrived at an embankment, apparently part of the old irrigation pond, when Carrie unexpectedly began hiking off the trail up the hill.

“What are you seeing?” she asked me, as I followed her onto the side of the embankment.

“Looks like some kind of layer of fine sandy stuff…” I responded hesitantly, “Oh, and the rocks are rounded.”

“You got it!” she proclaimed with a smile. “So, what we are seeing is lakebed and riverbed sands and cobble.”

Then turning, she pointed out to a suite of rimrock, lava plateaus.

“If you look across at our plateaus,” she explained, “you are looking at the old valley floors!”

She explained that each lava plateau was the result of an individual basalt eruption event (part of the Deschutes formation) that filled the valley at that point in time—the oldest being 7 million years old and the youngest only 3 million years.

Over time, the land area surrounding the lava-filled river channels eroded. As a result, what were once lowlands and river channels, are now basalt plateaus.

“This is inverted topography,” said Carrie—what was low is now high.

“What we are looking at here is the infill,” said Carrie looking back to the sand and cobbles, “the eroded remains of a valley bottom.”

Looking out at the lava plateaus

Perspectives

Carrie and I continue wrapping up and around the hill of infill where we could get a better view of the young lava flows and the much older rhyolite buttes of the Crooked River Caldera.

As we hiked, we passed by some bright yellow rabbitbrush still in bloom. Carrie told me how she uses it to make cloth dye; and we briefly got on a tangent regarding natural dyes—a side passion of Carrie’s.

“Rabbitbrush makes the best dye!” she proclaimed.

Speaking of color, Carrie pointed out a pale green patch of ground in the distance—to the left of Barnes Butte from where we stood.

She told me how she used to drive by and wondered at the green color—“it just stayed pistachio green” all year long. Eventually, she realized it was tuff.

Though the rock that makes up Barnes butte is a solid rhyolite dome, tuffs can be observed around Barnes Butte as a few outcroppings, and as what geologists call “float”—rocks that have moved from their place of origin.

Carrie pointed out a few outcroppings of Barnes Butte tuff that lay just in front of us—“the high points,” she noted. 

A Step Back

Carrie also addressed the hills that lay on the far horizon, outside the Caldera’s boundary.

“Most of what we are looking at on the far horizon are Clarno andesites,”  said Carrie looking east—volcanic rocks from a period preceding the Crooked River Caldera eruptions.

Of course, mixed up in all of it, is even older rocks. Accreted terranes—jumbles of earth materials that become permanently attached to a land mass of a completely different origin—make up the basement rocks of Oregon.

Carrie told me about how older maps used to show a pocket of limestone in the area. It was “weird” at first, but as Oregon’s geological story unfolded it became apparent that the limestone was from an accreted terrane. The limestone would have come from some distant shallow sea before it was added to the continent 100 to 400 million years ago by the forces of plate tectonics.

Only later it became part of the Crooked River Caldera. The past, literally, resurfacing by way of the Caldera’s eruption.

Flash Forward to Newberry

Carrie turned to face the interior of the Caldera again. There was still one more point in time to discuss.

In addition to the lava flows that make up many of the plateaus around Prineville, an even younger period of volcanic eruptions graced the Caldera in geologically recent times—the Newberry Volcanics.

Newberry has been erupting for the last 400,000 years and remains active today. Its most recent eruption was 1,300 years ago.

“Darn it all!” she exclaimed. “I was hoping it would be clearer…It [Newberry] is a big shield volcano,” said Carrie, “It barely shows over the horizon.”

Interestingly, some of Newberry’s flows reached into the Crooked River Caldera.

Carrie described one of these flows:

“That basalt flow was going down the ancestral Deschutes River, near O’Neil Junction, where it dropped into the Crooked River drainage, headed to Smith Rock. Here it smacked into Smith Rock pushing the Crooked River over to its present course.”

Those who have visited Smith Rock State Park and hiked any of its trails know this basalt flow as the calf-burning, heart-pumping climb out of the Crooked River Canyon, and back to the parking lot.

Next time you visit, “Look at what is at the bottom of the basalt flow…” advised Carrie. “There is river cobble there.”

Whether it is the Newberry basalt flow, or any one of the other flows that passed through, each time the Crooked River is displaced.

“It was doing its level best to be a valley bottom and these stupid basalt flows come in,” Carrie described in her own colorful way. “The river is like ‘okay, I will find another route’.”

Ashes to Ashes

At this point, Carrie and I resumed our walk along the old racetrack and took a left, wrapping around to the other side of the embankment facing Barnes Butte. Song birds flitted by as we walked. 

“One of the best-kept secrets,” Carrie shared, “we have a nesting osprey pair here.”

As we meandered around the bend, Carrie pointed out what looked like really fine sand.

“This is volcanic ash,” she explained. “When Mazama erupted, we got a foot and a half of fine ash.”

Mount Mazama—a massive stratovolcano blew it’s top 7,700 years ago, forming a smaller caldera that has since filled with water forming Crater Lake.

Carrie continued: “One of the things that happened is the winds will blow ash and it will catch on the leeward side of the hill,” she explained.

Carrie then proceeded to scoop up a handful of the ash and show how me how to look at it with a hand lens—white pumice fragments and black hornblende or magnetite could be made out among the grains. Of course, her favorite part, and mine too, was to look at the ash in the sunlight. 

“The best thing about volcanic ash is it winks at you,” said Carrie. “It is the reflection of the crystal fragment of volcanic ash.”

You don’t get that same winking with sand, explained Carrie. Only ash has the ability to sparkle.

Volcanic ash capable of winking in the sun

Blowing in the Wind

The ash is also important to the soil of the area. Loess—windblown sediment—is rich in many minerals and provides the starting material from which soil forms.

Of course, loess is not the only input into the area.

“Don’t forget we are in this pocket here,” reminded Carrie, “We had all the river systems and lake deposits that are actual sand and gravel.”

Alluvium—water-transported sediment—also contributes to soil formation, even in places you might not expect. Powell Butte, for example, is mostly covered with river sand.

“Something [i.e., a river] was moving across there at one time,” said Carrie.

Now, these old river channels are a ready source of water for the City of Prineville. When the City looked for places to tap for wells, surprisingly the best places were on the bottoms of the lava flows that once were river channels.

“This was the thing that blew me away,” Carrie smiled. 

Barnes Butte

Carrie and I reached another junction and took the trail heading up Barnes Butte. As we climbed, we passed by several large hunks of reddish-brown rock. Unlike the rocks down below, these were not round, but jagged.

“All the hunks of rock are rhyolite,” said Carrie.

I asked Carrie how she knew it was rhyolite, aside from knowing where we are at. Carrie picked up a piece of the rock and knocked it against another.

“It sounds glassy,” she explained. “Part is how it sounds, and if you can heft it.”

According to Carrie, compared to basalt, another prolific volcanic rock, rhyolite is not as heavy. So if you find a gray rock that is relatively lighter and glassier, it could be rhyolite.

Juniper

As we continued up the rocky hill Carrie, I noticed a juniper with its roots clinging to a juniper tree.

Off-hand I asked Carrie, “Do junipers like rhyolite?”

Surprisingly, she answered in the affirmative.

“That’s a cool story!” Carrie proclaimed. “Western Juniper has become invasive.”

Though western juniper is a species native to central Oregon, it has been creeping into areas that it normally wouldn’t. Fire exclusion, grazing pressure, and climate variability have all been cited as reasons for the spread of the waster juniper.

“And it uses a lot of water,” Carrie added, a highly valued resource in the area.

“This is all rangeland,” Carrie explained, it should have “more grasses and sagebrush component.”

In short, western juniper shouldn’t be so prevalent.

Instead, according to Carrie, western juniper is a first colonizer. Its range historically was limited to rocky areas—like our rock-grasping juniper.

“This is a rhyolite knob,” concluded Carrie, “and this is a very well-behaved juniper.”

Well-behaved journal growing from rhyolite rocks

Lichen

We continued up the Barnes Butte for a stretch but then decided to turn around. I was curious about finding tuff, so Carrie suggested we check the lower trail.

As we walked, I started noticing all the lichen and moss growing on the rhyolite and asked Carrie about it.

“Are they picky?” I asked, wondering if only certain lichen grow on certain kinds of rock.

Carrie didn’t think so, but instead mentioned how they might be used to age-date rocks.

Estimates of the age of a rock can be estimated based on the growth and size of the lichen that grows on it.

“Has the rock been sitting in place?” Carrie asked rhetorically. “Then you can get some age dates.”

Additionally, some plants do seem to prefer certain rock types. During the mapping of Mill Creek—an area adjacent to the Crooked River Caldera—McClaughry and Ferns found that, following a fire, much of the rhyolitic rocks were being colonized with manzanita. Manzanita soon became an indicator of rhyolite geology during the mapping.

Lichen growing on rhyolite

Recommendations

As we continued downhill, Carrie spotted some of the green tuff as float (loose rock) along the pathway—more evidence that we were, in fact, in a Caldera.

As we walked, Carrie offered me a lot of recommendations—video recommendations, places to visit, and hikes to take. She had a real knack for suggesting hikes I hadn’t been on.

But perhaps the strongest suggestion she has was to check out some of the Crooked River Caldera sites.

One of these places was Pilot Butte. (Yep, I hadn’t hiked it yet.)

You can see the Cascade Volcanoes from Pilot Butte—” a lovely white line of volcanoes,” as Carrie put it, but she wanted to make sure I didn’t miss the main event. 

“It [the Crooked River Caldera] is one huge volcano compared to the pretty pristine cones,” she added.

Other places she recommended for observing attributes of the Caldera include the Prineville Reservoir, Peter Skene Ogden State Park, Ochoco Reservoir, and, of course, Smith Rock.

I recommend hiking with Carrie. She is a hoot-and-a-half.


Carrie Gordon is a retired forest geologist. She was the Forest Geologist on the Ochoco National Forest and Crooked River National Grassland, U.S. Forest Service, headquartered in Prineville, OR. She retired in 2017. Carrie is also an active member of the Central Oregon Geoscience Society and an Oregon Master Naturalist through the OSU extension program. Carrie has had a life-long fascination with the land and the rocks, listening to the stories they tell.

Curious Hiker: John Dellenback Trail

Open dunescape on John Dellenback trail.

Overview

Hike through a unique dunescape, sculpted by the wind, on your way to the Pacific. The Oregon Coast Dunes National Recreation area is a place like no other—with a feeling of vastness, like the ocean itself. The John Dellenback trail is a favorite for hikers, as no motor vehicles are allowed in the vicinity. Forests, sand, and sea—enjoy the beauty and solitude this trail has to offer.

Highlights

Unique dune landscape; beach access; solitude/no vehicles; wildlife tracking.

Need to Know

Northwest Forest Pass or equivalent required at the trailhead. There is plenty of parking, and vault toilets available. No bikes are allowed on the trail. March 15-September 15 is western snowy plover nesting season. Trail users must remain on the trail and walk close to the water’s edge while on the beach during nesting season. No dogs are allowed during this time as well.

Hike Description

Forest Walk

The bridge that crosses Eel Creek.

A placard marks the start of the hike at the far north corner of the parking lot. From here, follow the trail across sandy-bottomed Eel creek on a wooden planked bridge and enter a forest of Shore pine (Pinus contorta) and Douglas-fir (Pseudotsuga menziesii).

Once an open sand dune itself, the forest is now thick with trees and shrubs that reach overhead. Showy, pink-flowered Pacific rhododendron (Rhododendron macrophyllum), thicket-forming salal (Gaulthoria shallon), and edible evergreen huckleberries (Vaccinium ovatum) proliferate here, as well as red-barked Columbia manzanita (Arctostaphylos columbiana).

Ignore a turn-off to the left, following a small wooden trail marker right.  Pass a wooden bench that looks out on a small pond.  Cross the campground road and re-enter the forest. A few lone madrones twist upwards amongst the conifers—red bark ablaze.

Dunes

Trail marker and dunescape.

Shortly, the path leaves the forest, spilling out onto open dunes. Mouth agape, begin heading west toward the ocean following a set of spaced-out wooden trail markers with a blue stripe near the top.

Here is where the fun begins! You can follow the marked route exactly or choose your own path through this surreal landscape. Sand shifts underfoot as you walk. Human and animal prints trail across the ground, impressions of the past.

Ripples on the sand.

To the left of the markers, a large oblique dune parallels the path. Oblique dunes are the largest dune type and can reach heights of 180 feet and lengths of up to a mile. Formed by the northwest winds of the summer and southwest winds of winter, these massive piles of sand are slanted and constantly shifting. Climb up the dune’s steep rippled side for a better view of the expansive sandscape.

Sideview of a large oblique dune.

Continuing west, notice a tree island to the left of the large dune and a low vegetated area to the right. Tree islands are remnants of past coastal forest that has since been buried beneath the sand.

Tree island to the left of the oblique dune.

Eventually, the sandy undulations head downhill toward a tree line in the distance. Mounds of sand with dune grasses—called hummocks—appear. Small water-filled depressions hide between the mounds, likely the result of a rising water table. Look for small plants like seashore lupine, beach strawberry, and dune tansy growing in the sand.

Deflation Plain

Entering the deflation plain forest.

Upon reaching a forest of shore pine, the trail takes a sharp turn to the right following the edge of the woods. Songbirds flit from branch to branch in this transition zone.

Again, the trail turns sharply, this time taking a left and diving into the forested deflation plain—a low wind stripped area full of wet sand and thick with vegetation. The area is swampy enough that a boardwalk covers a portion of the trail but be prepared to get your feet wet. Large puddles block the path frequently during the wet season.

Boardwalks help keep hikers’ feet dry.

Moss and lichen inhabit the sandy environment, and red bearberry or kinnikinnick (Arctostaphylos uva-ursi) grow in low dense patches—redberries stand out brightly against the narrow evergreen leaves. Shore pine trees crowd overhead, along with mossy Oregon myrtle trees in the swampy forest. Evergreen huckleberry, salal, and rhododendron reoccur here as well.

Foredune to Beach

Heading to the foredune and beach.

Eventually, the trail opens to a grassy foredune with occasional shrubs. Much of the grass that makes up the foredune is European marram grass (Ammophila arenaria)—brought to the coast intentionally to stabilize the otherwise dynamic environment. Fast-growing European marram grass has spread through much of Oregon’s central and south coast via a network of roots and rhizomes.

A small patch of European marram grass.

Dropdown off the grassy mounds of sand near a bright yellow sign marked 121—placed there for emergency responders, but also a handy way to remember the route back. Walk the beach for a while or simply enjoy the sound and sights of the rolling ocean waves before returning the way you came.

Time to walk the beach.

Optional:

After crossing the dunes on the way back, look for a small sandy trail that leads off to the right where you enter the forest. Follow this trail to hike the remainder of a small loop that will take you back to your car.

Hike with a Paleontologist

View into the badlands of the Blue Basin at the end of Island in Time Trail.

How would you like to travel back in time?

You might be thinking—sounds like science fiction.  And in a way, you would be right. But I am not talking about the sort of time travel involving flux capacitors or DeLorean Time Machines. Rather, the sort that uses dated rocks and hard scientific evidence to unlock the secrets of the Earth. 

Yes, I am talking about paleontology— a science dedicated to piecing together stories of the past. Stories of the evolution of a planet; stories of volcanism and climate change; of life adapting to land and to air; of speciation and mass extinctions. Just to name a few. 

With this in mind, I met up with Nick, Chief Paleontologist, John Day Fossil Beds National Monument, at the Blue Basin Trailhead in the Sheep Rock Unit of the Park, for a short hike and to listen to some of these stories.  

Hang onto your hats kiddos—we are headed back in time!

Nick Famoso next to one of three fossil replicas found on the Island in Time Trail

The Hike

  • Trailhead: Blue Basin Trailhead
  • Distance: 1.3 miles
  • Elevation Gain: about 220 feet
  • Details: There is ample parking at the trailhead and a pit toilet. Look for signs for the Island in Time Trail. There is also a 3+ miles Blue Basin Overlook trail you can take from the same location.

Rock Records

According to Nick, one of the main reasons why the John Day Fossil Beds National Monument was established was to “preserve and interpret the story of the geological past of the John Day region.” The park represents over 40 million years of time. With the Sheep Rock Unit providing the longest geological record from about 33 million to 7 million years ago. 

Nick and I met up to hike the Island in Time Trail, which takes you into a small section of the Sheep Rock Unit known as the Blue Basin and/or the Turtle Cove assemblage. The fossils found in the Blue Basin represent a relatively small slice of geological time, from 30 to 29 million years ago, but a well-preserved slice. Within this small slice, each layer of ash can be dated to about 10,000 years. As Nick explained—”the ashes are like page numbers” allowing for very precise (geologically speaking) dating of fossils found in the rocks.

Looking down the trail into the badlands.

Condon 

Nick and I started down the trail but quickly stopped short. Before going back in time millions of years, we needed to go back just 150 to meet a man named Thomas Condon. 

As Nick relayed the story:

Condon was born in Ireland before moving to New York as an older child. He always had an interest in rocks and fossils. He even had a small collection that he brought with him to the states.  As an adult, Condon’s fascination grew. Even after becoming a reverend and moving to Oregon, he continued to collect and discuss his fossils.

Eventually, Condon set up his ministry in The Dalles, Oregon where he developed a reputation—spreading, not only his religion, but his passion for geology. His reputation grew to the point that visitors, usually soldiers, that passed through would bring him fossils collected during their travels.  Many of these fossils came from the John Day region. 

Intrigued by these gifts, in 1865, Condon decided to visit the John Day region himself. A visit that would change his life and the face of paleontology in Oregon forever.

From that point on, Condon traveled and collected fossils all over Oregon, adding to his personal collection along the way. Eventually, Condon would become Oregon’s first state geologist and the first natural history faculty member at the University of Oregon, where his fossils still make up their core collection. 

Nick explained that Condon would have specifically visited the Blue Basin. “What Blue Basin is really interesting for is getting directly into the rock layers Condon would have been working with.”

Fossil Land

Following in Condon’s footsteps, Nick and I resumed our hike. But before long something caught his eye.

Nick pointed to a bright white patch of ashy looking material in the rock layers. “When you see pockets like that,” he said, “that are bright white, it is most likely Mount Mazama ash.”  Mount Mazama erupted 7,700 years ago, dropping several inches of ash over much of the Pacific Northwest, leaving a massive depression that would eventually become Crater Lake.

I asked Nick, why the Mazama layer could only be seen in this small pocket and not as a continuous layer within the earth’s crust. He explained, in a perfect scenario, Earth material is laid down in even horizontal layers under the action of gravity—what is known as original horizontality. However, most landscapes are not flat, and Earth material doesn’t settle in place. Instead, paleotopography—the shape of the land in the past—shifts material, changing where and how much is deposited across the landscape. For example, you might find an eight-foot ash deposit in one place (perhaps a basin of some sort), while in another place the deposit may be only two feet deep, or not exist at all.

These basic principles also help explain why fossils are only found in certain places on Earth. As the sediments involved in fossil bed formation are also influenced by paleotopography.

Floodplain Deposits

However, the shape of the land is not the only important factor for fossil bed formation. The movement of water and wind is equally important. As agents of erosion and deposition, water and wind influence where sediment accumulates, potentially burying the remains of organisms and creating the conditions for fossil formation.

The fossil beds at Blue Basin formed in a river valley that experienced frequent flooding. With each flood, sediment was deposited on the banks, and in the floodplains of the river, burying animal remains in its wake. Overtime, sandstone, siltstone, claystone, and other sedimentary rocks formed, preserving hard materials, like bone, that would fossilize.

The white layer seen here is probably ash from the Mount Mazama eruption of 7,700 years ago.

Colorful Rocks

Moving deeper down the trail, Nick and I entered an otherworldly basin of pale-colored rock. The walls of the basin, like a layered cake—with each layer a different color and thickness. 

Nick explained how each layer formed. Blue-green colored layers are attributed to the mineral celadonite— formed during diagenesis—the conversion of sediment to rock. Brown colored layers are claystones that didn’t undergo a secondary alteration process. Bright white ledges are volcanic tuff —formed from the compaction and cementation of volcanic ash. Embedded in the rock layers were also concretions—compact masses of sedimentary rock that form around a nucleation site, like a bone. Fossils can sometimes be found encased in concretions. 

Ultimately, the stratigraphy of rocks helps paleontologists establish boundaries between subunits of rock—each of which represents slightly different environments.  In Blue Basin, there are 7 subunits that have been characterized by their geological composition and ashes—B through F with some letters divided further. 

Can you see some of the many layers of rock in this exposure?

A Layer of Ignimbrite 

Nick also pointed out a thick brown layer of rock that capped a section of the Blue Basin rocks. This layer is visible at various places within the Sheep Rock Unit, as well as other areas in the Park. 

The layer is ignimbrite, rock formed by the consolidation of pyroclastics, or as Nick put it—”a fiery cloud of death.” These “death cloud rocks” were the result of eruptions from the Crooked River Caldera—a now extinct supervolcano that once covered several square miles in central Oregon. 

Fueled by an early version of the Yellowstone hotspot, The Crooked River Caldera eruptions would have been huge—on a scale far beyond what we have seen in human history. For his dissertation, Nick studied the impacts and recovery of life following ignimbrite eruptions, finding that these sorts of large impacts seemed to have equally large, long term impacts over time.  In the case of the Picture Gorge ignimbrite, there is evidence that the landscape changed from more forested to less forested due to the eruptions. 

Hidden in the Rocks

According to Nick, stories of change are what make paleontology such an important field of study. Stories about changing climates, mass extinctions, and catastrophic events are all hidden in the rocks. If we pay attention to these stories they can help us gain perspective on issues we face today.  

For example, scientists agree that we are in the middle of a sixth mass extinction brought on by human activity.  But because it is occurring on a time scale that is outside human experience—thousands to millions of years—we don’t see it. Paleontology can help us understand this discrepancy, giving us the opportunity to respond accordingly. 

Collecting Fossils

Nick and I hiked deeper into the badlands of the Blue Basin. Nick said that on any given day, approximately 10 field collections might be extracted from this unit. Considering that the Blue Basin for over 30 years, that is a lot of fossils!

I asked Nick, what exactly constitutes “a collection?” He responded, “Usually something that can be identified to a fairly high level.”  For example, identifying a tooth as Mammalian would not constitute a collection. But identifying a tooth is from a rhino, well then you have something! Teeth, bones, seeds, tracks, and/or traces of past life could potentially end up in a collection.

However, no fossil is collected alone. “What is most important is the context,” explained Nick. Gathering material and carefully documenting where a fossil is found is often more important than the fossil itself.

Any specimen found loose, or “in float” is put into a bag with any other material that is found within a three-meter area. Fossils that are found “in situ,” or in the rock, also require detailed documentation of fossil location and position in the rock, as well as other contextual info. Either way, the more contextual information gathered, the better! 

Past Life

“Fossils are evidence of past life.” So, what life existed in Oregon’s John Day region about 29 million years ago? 

Well first, picture a river valley; open and expansive with rolling hills and dales. The climate would have been dry and cool—suitable for the hardwood forests and open meadows that permeated the landscape at the time.

And as for the animals—there were a lot of them!  According to Nick, the diversity of life that once existed in the John Day region was tremendous—with at least 100 different extinct species of vertebrate life has been found in the Turtle Cove assemblage.

Herbivores

Most abundant were herbivores, specifical ruminants like Hypertragulus—a mouse-deer creature—which make up about 47% of fossils collected in the Turtle Cove Member. There were also Oreodonts—large sheep-like and pig-like even-toed ungulates of which there are no modern descendants. As well as one to two species of horse, like Miohippus, a three-toed horse.  And, for good measure, large rhinos roamed the valley. 

Carnivores

Then there were the carnivores.  Though not as abundant, the diversity of carnivores that existed 29 million years ago is impressive. At any given time, there would have been up to ten species of dogs, like Mesocyon, coexisting together by taking on unique roles in the ecosystem.

In contrast, today there is only one living species of dog in the world. The rest have gone extinct as other groups of organisms, like weasels, came onto the scene. Nick said that this evolutionary see-saw is pretty typical of life on the planet. “Depending on what is going on and the evolutionary process,” the pendulum swings and different groups of organisms become more dominant.

Nick was also quick to point out that this doesn’t mean that one group of organisms is better in some way than another. “No animal or plant that is alive today is no more or less evolved than anything else alive today,” said Nick. “Fish used to be the most dominant things,” explained Nick, then reptiles. Mammals and birds came later, but they aren’t more evolved.  If you want to get really technical, you might say all vertebrates are really just a bunch of fish—we are all equally evolved from a common fish-like ancestor. Just because some species don’t look much like fish anymore, doesn’t make them better. 

A Word On Plant Fossils

There were of course a lot of plants in existence 29 million years ago, but you won’t find a lot of plant fossils in the Turtle Cove collections. Why? Well for one, in general, plant fossils are harder to find.  You might get a part of a plant, like some wood or a leaf or root fossil, but rarely the entire plant fossil. This makes it difficult to get down to a species level of identification, explained Nick. More often paleontologists must be content with community-level identification of plants.

In addition, the conditions required for fossil formation is different for different forms of life. In the Blue Basin, the conditions were likely too destructive for plant fossils to really form. Though there are a few leaf and seed fossils found in the area, most plants would have probably washed away or been eaten during the flooding events that laid down so many vertebrate fossils. Old lake beds are often great places to find plant fossils, explained Nick, because the environment is calmer.  Plant fossil formation also depends on the acidity of the environment as well.

Replicate

One group of charismatic animals captured in the Turtle Cove rock layers are multiple species of a land tortoise—Stylemys. In fact, you can visit Stylemys on the trail! O.K., well not actual Stylemys, but a replica of a fossilized male. 

When we reached the replica, Nick and I stopped to chat.

There is a lot you can learn about an animal from a fossil. For example, in the case of our land tortoise, the plastron—the protective front of a turtle, opposite the shell—can tell you if the turtle is male or female. Nick explained that males have concave plastrons so that when they are mating with a female they do not roll off her.  This is true of turtles alive today. Thus this trait has been around a really long time, providing a direct link between the past and present.  From an evolutionary standpoint, this makes sense—a trait so vital to reproduction is bound to provide an evolutionary advantage. The survival of the fittest only works if you can “replicate.”

A fossil replica of a land tortoise as displayed on the trail.

Replicate

Nick and I continued down the trail until we reached another fossil replica. This time it was of a sheep-like Oreodont.  A browser, the fossil showed pointed canines, used for snapping branches, and a large depression above the cheekbone, indicating huge strong chewing muscles. On the face there were also two other small rounded depressions—one for the eye and the other a scent gland that can be seen in even-toed ungulates today. Though it is unclear what the gland was used for by Oreodonts, one might predict its function was somehow beneficial to the organism—perhaps used in sexual selection.

However, though millions of years ago there were dozen of species of Oreodonts in North American, none exist today.  A placard found next to the fossil asks, “An evolutionary success?”

It is really easy to dismiss extinct species. But Oreodonts survived in North American forests for over 30 million years, much longer than most animals alive today. So as we learn about their loss, perhaps there is an even greater lesson to be taken from Oreodont success?

A fossil replica of an Oreodont as displayed on the trail.

Small Things

As Nick and I stood by the Oreodont fossil, he also pointed out a layer of sandstone in front of us. He told me that this layer wasn’t the usual overbank deposits found throughout the basin, but an actual in channel or river deposit. 

The great thing about river deposits like this one is that they “produce a lot of smaller fossils…and smaller things tell us a lot more about the environment than bigger things,” said Nick. Because small mammals usually have smaller ranges, they provide information about local conditions. The populations of rodents living in the Blue Basin stayed in the basin and were endemic to the area. Rodent populations in other areas would also be uniquely adapted to their environments. 

Some small species can even act as geological time markers and can help paleontologists understand what is happening on a global level. Nick mentioned the Hypertragulus as an example. In the Great Plains, the Hypertragulus went extinct, while persisting several million more years in the west. By tracking differences in populations this way, regional stories unfold as scientists question the discrepancies. It raises the age-old question: Why?

A river channel deposit of sandstone.

The Species Problem

Before we moved on, Nick also brought up an important paleontological problem—the species problem.  In general, scientists define species of genetically similar organisms that can exchange genetics and/or interbreed. But, as Nick put it bluntly, “you can’t exactly put two fossils in a box and see if they make more fossils.” 

Therefore, defining a fossil species is complicated. In the past, species were defined based on theoretical ideas, like the likelihood of geographic separation between similar-looking fossil organisms. However, this is not a very reliable way of distinguishing between fossil species. There is a lot of natural variation within populations. Thus, simply finding small differences between two groups of similar organisms does not a species make.  

Nick shared an example of the species problem he encountered in his own work. He had looked at the differences between fossils from eight different species of horse. He compared the variation he found between these fossils to the variation found in modern horse and tapir species. The differences were not significant. Nick concluded that eight fossil species of horse were better grouped as two.

The Specialist Problem

Nick and I continued down the trail until we reached a final fossil replica. The replica was of a false saber-toothed cat— a nimravid. Cousins of true cats, nimravids branched out from other carnivores about 16 million years ago.

During the time period associated with the Turtle Cove assemblage, only three or four nimravids coexisted.  They varied in size, with false-saber toothed cats the largest of the carnivores overall. 

However, despite their size, nimravids were more prone to extinction. You see, nimravids tended to be specialized meat-eaters. Meaning they relying on only certain food sources for survival.  In a stable, unchanging environment, this is not particularly problematic. It is actually a good way for species to carve out a role in the web of life.  

However, environments change and this is where being a specialist can be a huge disadvantage.  It is sort of like the old adage—don’t put all your eggs in one basket.  If that basket gets knocked over, no more eggs. 

Generalists, like many dogs, on the other hand, tend to do better, as they rely on a varied diet and lifestyle—they have their eggs in many different baskets. 

Of course, Nick added, being a bone crusher is one specialization that has been successful. There always seems to be enough bones. 

Amphitheater 

Eventually, Nick and I reached the very end of the hike—a place known as the amphitheater. Here it is easy to see the many layers of rock.  The “pages of time” literally surround you. Nick called out each layer: lower green is unit C, browns unit D, the ledgy layers E1-E3, followed by the Blue Basin Tuff and unit F, with dark Picture Gorge ignimbrite capping it all.

Layers and layers of rock containing fossils of past life, telling a 29 million-year-old story. 

The amphitheater is the end of the trail.

Storytelling

As we made our way back out to the trailhead, Nick and I continued to discuss what it is like to be a paleontologist. At one point, I asked him what he liked most about the work.  It wasn’t the story making, but the storytelling—the advocating for the fossil resources of the park that he felt was most important. 

It is the work of the paleontologist, not only to piece together these stories but also to tell them. As Nick said earlier during our hike: “I tell students all the time that the most important thing, as a scientist, is that you have to be able to communicate what you have done to somebody else.  Because if you can’t, what is the point of doing it in the first place?”

I couldn’t agree more, Nick.

Nicholas Famoso (Nick) is the Chief Paleontologist and Museum Curator for John Day Fossil Beds National Monument. Nick got his bachelor’s degree from South Dakota School of Mines and Technology where he studied fossil mammals and marine reptiles. He later went on to earn his Masters and Ph.D. from the University of Oregon in geological and earth sciences.

Hike with a Botanist

View of Iron Mountain from Cone Peak.

I don’t know what it is, but I love plants! You know, the ubiquitous, but easily overlooked green stuff. When plants arrived on the scene they literally changed the world! Nearly all life on the planet depends on plants for survival. And they are pretty, oh so pretty, What is not to love? 

Visions of wildflowers danced in my head as I drove out to meet with Linda Hardison, Botanist, and director of OregonFlora, at the trailhead for Iron Mountain and Cone Peak.  It was July 2nd—the perfect time of year to catch the wildflower show the area is known for.  Set aside as a Special Interest Area for botany, Iron Mountain and Cone Peak attract the attention of many adventurers looking for botanical inspiration. 

So I guess I shouldn’t have been surprised when I pulled into the parking lot at Tombstone Pass to find it rather full for a Thursday morning. Unfazed, with sunscreen, hat, and face mask in hand, I found Linda at the other end of the parking lot.  We quickly exchanged greetings, before hitting the trail.  

All set for a riotous romp filled with botanical delights, I was not disappointed.

Linda Hardison on the trail.

The Hike

  • Trailhead: Tombstone Pass Trailhead
  • Distance: 7+ miles
  • Elevation gain: approx. 1700 ft
  • Details: Amply parking and pit toilets at the trailhead. Trailhead is very accessible as it is right off Highway 20. You will need to cross the highway during the hike. Trails are well marked and maintained.

Plants Rule

Linda and I began our hike from the tombstone pass parking lot, heading down the trail toward Cone Peak. As we ducked down into the foliage and made our way along the trail that leads through some wet meadows, I immediately peppered Linda with questions: Why plants? Why study Botany? Why is botany important? 

O.K. so I was a bit excited. Linda was gracious with her reply.

“Well, just look around—it is just so beautiful,” she said. 

But in true scientist fashion, she elaborated, “Botany is at the core and foundation of everything. It’s what lets it all happen.” She went onto explain how flowers evolved this amazing ability to do photosynthesis—they capture sunlight and store it in organic molecules. These organic molecules are the basis of the entire food chain, feeding other living things, as well as enriching the soil for more plants to grow in. 

“The whole planet depends on plants,” she stated. From a humanistic standpoint, we need them to breathe, eat, and build with. They are so fundamental and they are everywhere. 

Plants are everywhere

This is the other amazing thing about plants! They are literally everywhere. “Plants have adapted to every condition on the planet,” explained Linda. Over millions of years, more and more species of plants have evolved and taken on different ecological niches, or roles, in the environment.  We now have entire communities of plants that live in forests; others in meadows; some grow in valleys, and others on mountains. Each plant with its own way of surviving in these different conditions. 

A Flora 

A whole planet to cover, there are a lot of different species of plants on Earth.  Too many for one blog post to focus on, but if we narrow it down to say— Oregon—the task goes from being impossible to overwhelmingly difficult  O.K. still too much for a blog post, but not too much for Linda and her Colleagues who are taking on just that within the OregonFlora program.

Linda shared some background on the project. The Oregon Flora Project was founded in 1994 by botanist Scott Sundberg with the goal of creating a new flora—basically a plant identification and information manual—for the state of Oregon. At the time the most up to date manual on Oregon plants was nearly 50 years old, so it seemed like an update was in order.  As Linda explained how plant populations change: “new plants are always being discovered; new species come in either as weeds or as climate changes and new habitats open up and they adapt to new places, and things go away, things get extirpated.”   

So with access to Oregon State University’s herbarium the project got underway. Currently, Volumes 1 is available and Volumes 2 & 3 are in the works.  OregonFlora also has a website chock full of botanical information (a massive overhaul is underway to make it more user friendly) and an application—Oregon Wildflowers App—that I personally love and use to identify plants while hiking. 

As for the plants themselves, OregonFlora has captured information on 4,762 different plants in Oregon. Talk about biodiversity!  

Ecoregions 

The diversity of plants in the state makes a lot of sense when you think about the diversity of ecosystems and habitats that exist in Oregon.  The state of Oregon is somewhat unusual in that it has many different ecoregions—large areas of the state with similar climate and vegetation. From the cool, wet Oregon Coast Range to the hot, dry Basin and Range—Oregon has been cut to shreds ecologically. 

OregonFlora uses the ecoregion concept as a frame of reference as well—recognizing 11 different ecosystems that largely parallel the EPA designated ecoregions in Oregon. Plant communities can be organized within this framework and then further subdivided further into habitat preferences from there. 

Linda told me that when she gives presentations about OregonFlora, she will put up a picture of open scrub and grasses and ask people where the picture was taken? “And people will say eastern Oregon,” said Linda. Then she will put another picture up and the audience is right again. “I think people really have a gestalt about this sort of stuff,” explained Linda. People recognize the combination of physical and biological attributes that make up a place. And OregonFlora is there to help them simply hone their awareness.

What’s in a name?

As the trail really started heading uphill, Linda brought up another botanical topic of importance—naming.  One of the big jobs of OregonFlora, and a big part of botany as a whole, is figuring out what to call a species. Naming is important because 1) they allow scientists to be precise when they are talking about plants, and 2) because they reflect the evolutionary relationship between plants. As simple as this sounds things can get a bit complicated. 

During the 18th century, Carl Linnaeus, often considered the father of taxonomy, first came up with the binomial, two name system that is, for the most part, used today. According to Linda, his system grouped organisms by morphological characteristics, especially reproductive features. At the time, these features were considered immutable because they were so important to the survival of a species. However, with the advent of DNA sequencing, we now have more complete information about how organisms are related.  We can see in the DNA evolutionary relationships that don’t necessarily match up with the optics.  

So names have changed. And scientists have had to adjust. “Botanists can take care of this,” said Linda by using synonyms. Botanists keep track of and acknowledge synonyms—old, out of date scientific names—as part of their records of each species. 

Linda pointed out some false Solomon’s seal. Once Smilacina racemosa it is now part of the maianthemum genus with the scientific name: Maianthemum racemosum

False Solomon’s seal (Maianthemum racemosum)

Not so Common

Then of course there are common names, like false Solomon’s seal.  OregonFlora also keeps track of common names.  Linda said they look for the most widespread names for each species in the region to include in the flora. Choosing regionally significant names of species is important as there can be great variability region to region.  For example, OregonFlora’s emblem is what most Oregonians call a fawn lily, but in other parts of the country, the same species is known as a dogtooth violet. Though these names don’t provide precise information about the plant, they do provide regional and cultural context to the botanical world.

Taxonomic Concept

But the division of plants doesn’t start and end with a name (or two or three).  As Linda puts it, “the isness of a plant” must be determined. The isness of a plant—its characteristic features and range—is what is known as a taxonomic concept. So considering our false Solomon’s seal the taxonomic concepts answer the questions- What is a false Solomon’s seal? It breaks it down and “puts a circle around it.” A name doesn’t mean a whole lot if it doesn’t have a specific plant associated with it. 

Taxonomic concepts might be species specific or subdivided further into varieties or subspecies. Linda said, “We try and get down to as small a bucket as we can.” We talked about the example of the species Pinus contorta.  The variety of this species that grows on the coast takes on a twisted, gnarled shape, but the variety that grows in montane environments is narrow and upright. 

Determining a taxonomical concept for a plant is challenging because there is natural variation in plants as well.  Genetic biodiversity that makes one member of a species different from another member of the same species can sometimes blur the lines and create some controversy or debate. Also, things are always in flux.  Whether it is through scientific work or natural processes, taxonomic concepts can and do change. 

Forest Habitat

By this point, Linda and I had hiked up out of the meadow and into the forest. Conifers dominated the overstory— Douglas-fir and hemlock primarily, but also some true fir and western red cedar. The understory was shaded, and the soil rich in organic matter and moisture captured by the trees and the fungal network below. Plants like vanilla leaf, wild rose, wild ginger, thimbleberry, bunchberry, and vine maple, took up residence in this protected understory. 

It is the combination of physical and biological features that create “the magic mix” for a habitat, explained Linda.  You can’t rely on physical features alone to determine what lives where.  As important as factors such as light, temperature, and moisture are, the other plants, animals, and microbes that live in the same community are just as (if not more) important.  That is why you can travel to places that share physical features, Linda explained, and find a completely different suite of organisms. 

Forest habitat from the trail.

It takes a Community

I asked Linda to elaborate on the importance of community connections. She was clear—we still have a lot to learn when it comes to relationships between species. We have an inkling that these connections are important, but “we don’t know what makes everything work together,” she said.

In terms of the forest, for example, we are just starting to learn about the importance of fungal communities in communication and the exchange of nutrients. Linda said that she knows a landscaper who will save the “duff layer” of an area before it is bulldozed for development so he can reuse the material later. 

Natural and wild places are just that important. “We don’t know what sort of glue some tiny little scrubby looking plant might serve in the big picture,” said Linda. “If we view things as a system, as a whole, I think there are much richer opportunities to learn and explore and benefit from than from looking at one isolated species.” 

Botany Rocks

Eventually, Linda and I stepped out of the trees and out onto a rocky outcrop to be greeted by a gorgeous wildflower display—the first of many. Bright yellow Oregon sunshine, deep purple larkspur, and bright blue gilia were all at peak bloom! We gawked at the beauty of the place—bright spots of color and fragrant smells overwhelmed our senses. 

First rocky outcrop we saw on the trail.

Adapt 

However, perhaps even more fascinating, is that these plants exist here at all. Living in a harsh environment with little moisture, very little soil, and a lot of snow and sun—these plants were not here for our benefit, but because they had adapted over generations of time to these conditions. 

Adaptations are characteristics that allow a species to survive and reproduce in their environment. Adaptations arise over multiple generations of time through the process of evolution by natural selection. Though it is impossible to observe this process on a hike, adaptations are readily observable. 

Linda and I speculated on some of the adaptations observable on our rocky outcrop. We noted how stonecrop and a species of claytonia both had fat fleshy leaves used to retain moisture. Cat’s ear lily uses a bulb for storing resources for the long winter. Blue gilia was prolific—adapted by living an annual lifestyle—producing a lot of seeds before dying off. Rough paintbrush, a hemiparasite, takes advantage of the company of others, stealing water and nutrients from their neighbors. Then there were the bright colors and fragrant smells of many of the flowers—all adaptations for attracting pollinators quickly during a short growing season.  

Blue gilia growing on the rocky outcrop along with larkspur and yellow monkey family.

Family Ties

As you can imagine, walking among the rocky outcrops the diversity of plants was captivating. We continued along the cone peak trail, stopping to admire, identify, and take photos along the way. 

At one point while discussing a species of buckwheat (Polygonaceae), Linda shared with me her secret to identifying plants. She said that one of the best ways to learn (for her anyway) is to look at the relationships—the family of plants. “It gives you the start on what something is,” she said. “If you learn the characteristics of a family, it opens the first door.” Then you can use a field guide or an app to narrow things down. 

The buckwheat family, for example, typically has lots of very small flowers and colored or no sepals. Asters (Asteraceae)—the largest family—tend to look like sunflowers with very open blooms that are attractive to pollinators. Oregon sunshine is a great example. Some relationships are surprising and a little more challenging. For example, Larkspur is in a subclass of the buttercup family (Ranunculaceae).  Characteristic of this subgroup are the fruits or “follicles”— capsules that open along a single side. In general, one way to recognize buttercups is to look at the fruit. 

Arrowleaf buckwheat (buckwheat family), Oregon sunshine (aster family), and purple larkspur (buttercup family).

Botanical Controversy 

One “family” of plants that I personally enjoy are the penstemons. I asked Linda what characteristics are common among penstemon.  She told me that they typically have “snapdragon looking flowers” and that the common name for the family actually used to be snapdragon (Scrophulariaceae). However, according to Linda, the “scroph” family is a classic example of “plant families gone amuk.”  As botanists have grown to better understand evolutionary relationships, the family has actually been split into five different plant families!  According to Linda, this change sparked a lot of frustration, as people had been trained (like herself) to identify these plants as “scrophs.” She admitted that she still often gets these plants mixed up. Botany is not without its challenges. 

Deadly Problem

Identifying plants is not only a lot of fun, but really important if you spend any amount of time outside. Plants can act as irritants and toxins.  Most people are aware of plants like poison oak and poison ivy, but there are many other plants that can cause problems. Giant hogweed, for example, can induce photosensitivity in people that touch it.  

On our hike, Linda and I ran across several stands of death camas. Death camas is lily-like in the false-hellebore family (Melanthiaceae)—characterized by parts in groups of three—but many people confuse it with real camas, also a lily, and an edible plant. The problem arises from the fact that they can grow in similar environments and are harvested as bulbs that can be hard to tell apart. Being attuned to the differences between these two plants is literally life or death. “Plant families gone amuck”

Stand of mostly death camas with larkspur and paintbrush.

Microhabitat

As we finally reached the sloping top of Cone Peak, Linda and I noticed some areas of land that looked a bit different from the rest of the bloom area. The mix of species was a bit different, with certain species more abundant, while others less abundant—more moss and grass especially. Even the physical characteristics looked different—more rocky and dry; the site also looked like it might have been more recently disturbed.  

Whatever the specific reasons, this was an excellent example of a microhabitat. Even within a defined habitat, there is small scale variability that can alter plant communities. According to Linda, understanding microhabitat is really critical for planting projects and restoration work. It is also something we still don’t know a lot about.

Honing our awareness for microhabitats is also a fun way to think about botany while on a trail, with the potential to contribute to a collective body of botanical knowledge. 

Unusual section of Cone Peak bloom— rocky microhabitat.

Drawing Connections

Speaking of fun ways to interact with plants, one of the goals of OregonFlora is to encourage people to engage with botany.  This is also one of my goals in writing this post.  There are, of course, a lot of ways to do so (some already mentioned in this post), but as we hiked from Cone Peak to the Iron Mountain trail junction, I asked Linda what she thought someone might do to develop a botanical eye. 

Linda’s advice was to first—“stop and look.” And second—draw! Drawing is a great way to pay attention and notice details that you might not notice from a glance or even a picture.  

Take some time to stop and sit down with a plant, suggested Linda. Pay attention to the sites and sounds. Look at what is covering that ground and what makes up the overstory.  “Understand this is a community: and who is a part of it, and who the big players are and who are the quiet voices,” said Linda. 

More blooms on Cone Peak.

Ethnobotany 

At one point, as we headed down the trail, Linda noticed a nondescript plant, a biscuitroot (Lomatium). She pointed out that indigenous people in Oregon often collected biscuitroot tubers as a food source.  The ethnobotanical aspect (or traditional use) of plants is another way people can relate to plants, Linda surmised. 

The Summit

Finally, we reached the trail junction for the iron mountain summit and we decided to make the ascent.  And after huffing and puffing our way by many more wildflowers, we reached the top and some amazing views of the Cascade Peaks. 

But the mountains were not the only thing in view.  Linda pointed out, as we looked out across the landscape, the diversity of the plant life that, though we couldn’t see the details of, blanketed all surfaces.  Old and new forests, open meadows, riparian corridors, and landslides—the view was awash in greenery.  

View from the summit.

As we made our way back down, more quickly than we went up, we continued to chat about plants and education, equity, and web design, among other topics. Our conversation shifted from topic to topic, almost as quickly as the plant communities changed as we moved down the mountainside—a dizzying array of botany.

So what is the takeaway? What did I learn from my hike with a botanist? Well to sum up: we are just a single species living in a world dominated by plants. So, as the saying goes, take time to stop and smell the roses.

Linda Hardison is a research assistant professor in the Department of Botany and Plant Pathology at Oregon State University and is the director of OregonFlora. She received undergraduate degrees in botany and marine biology from the University of Texas, and a Ph.D. in botany from the University of Washington. She currently serves on the board of the Native Plant Society of Oregon.

Resources

  • http://www.oregonflora.org
  • Meyers, S.C., T. Jaster, K.E. Mitchell & L.K. Hardison. 2015. Flora of Oregon. Volume 1: Pteridophytes, Gymnosperms, and Monocots. Botanical Research Institute of Texas, Fort Worth, TX.

Hike with a Habitat Wildlife Biologist

View of the Sandy River from the trail at Oxbow Regional Park.

10 Essentials

When you are out hiking, it is often recommended that you bring along “the 10 essentials:” navigation, sunscreen, knife, first aid, headlamp, fire starter, shelter, extra clothes, extra food, and extra water. These items are necessary for your survival, especially when things don’t go according to plan. 

Wild animals also have “essentials”—things they need to survive. However, unlike humans, they can’t carry these in a pack but must find what they need in their environment. In a healthy, unaltered ecosystem, this can be a challenge. In a heavily impacted ecosystem, it can become impossible. 

Meet Bill

As a habitat wildlife biologist for over 40 years, Bill Wieler’s CV is jam-packed with conservation, education, and restoration work. Bill has spent his entire career studying how to best protect wildlife and ensure their essential needs are met; as well as, worked on countless restoration and enhancement projects to that end. 

So when I met up with Bill at the Alder Group Picnic area at Oxbow Regional Park near Gresham, Oregon to begin our hike, I was thrilled to learn more, not just about wild animals, but the places they live and how we can do our part to protect them. 

Bill Wieler standing next to a Pacific yew.

The Hike

  • Trailhead: Alder Group Picnic Area
  • Distance: 2+ miles 
  • Elevation Gain: approximately 400 ft
  • Details: It costs $5 for parking. There are about 12 miles of hiking trails to explore. There are many different trail entry points to choose from. 

Yew Should Consider the Yew

Starting down the trail, one of the first things Bill pointed out was a Pacific yew tree. I love Pacific yew trees. As slow-growing conifers, they are often hidden among taller, more conspicuous trees. I often find them dripping with layers of moss and lichen, almost entirely concealing their noteworthy beautiful red bark. 

“It used to be considered a weed tree with no timber value,” said Bill, referring to the yew. He explained, only later, did scientists discover that its bark could be used to produce a cancer-fighting drug called taxol. “I always mention yew because it shows we really should be taking care of everything because we don’t know what animals and plants can provide.” 

Pacific yew along the trail.

Wildlife is Essential 

Wildlife is essential—it provides a host of benefits. Not every species will provide a cancer-fighting chemical like the Pacific yew, but ecosystems that contain a lot of different species have the potential to provide a myriad of benefits. 

According to Bill, “we depend upon natural ecosystems for many of our needs.”  “Food, fuel, and fiber” are perhaps the first of these benefits that come to mind. However, there are many less obvious benefits, including clean air and water, decomposition of wastes, and flood protection. This is not to mention the many social and emotional benefits biodiverse ecosystems offer. As Bill puts it, “they give us a complete, more healthy world. They enhance our sense of wonder and place.” 

Edible red huckleberries were abundant along the trail.

Moving Up

Unfortunately, most species are lost quietly without anyone noticing—species are lost before we even have a chance to appreciate their value. Even well-known species have faced threats because we have undervalued them.  For example, during the 1900s wolves were nearly eradicated from the lower 48 states in the U.S. because they were seen as dangerous to livestock operations.

Bill is optimistic though. He told me how he often polls people regarding their feelings on various wildlife species. And wolves, among other once-hated-species, have been moving up the list. As scientists have come to recognize the role of top predators in keeping other populations in check—what Bill referred to as “trickle-down ecology”—public acceptance of wolves has improved. For some reason, Bill hasn’t seen a large shift in public opinion when it comes to mosquitos and ticks though. 

Forest Dynamics

As we hiked deeper into the Douglas-fir/Hemlock forest, our conversation shifted from individual species of trees to consider forests. Forests are more than a collection of trees. Rather, healthy forests are dynamic ecosystems that operate as a unit. In fact, many of the wildlife benefits discussed earlier are really a function of a biodiverse ecosystem and not individual species.

According to Bill, there are six structures of a healthy forest ecosystem: 1) big trees, 2) snags, 3) logs, 4) soil, 5) open spaces, and 6) canopy layers. All of these components work together to keep the ecosystem functioning. 

While big trees provide excellent habitat for some species, like bats; when trees fall in the forest, it opens up space for new species and canopy layers to grow. These new species provide new resources and increased ecosystem resiliency. In addition, the down-wood and standing dead trees called snags, that remain following a blowdown, provide habitat for a host of insects and bacteria.  In fact, dead trees can host far more species than living trees, according to Bill. Then, over time the woody material decomposes, which builds the soil, providing nutrients for the next generation of forest plants. 

Can you find the six structures of a healthy forest in this picture?

Observations of a Forest

Bill pointed out that all of the six structures are observable in the forest at Oxbow Regional Park. The park even has some old-growth forest within its boundaries. In contrast, a forest that has been managed for timber production is less likely to contain all of these structures and/or in less abundance. For example, 8-10 snags per acre are typical of a healthy forest, while forestry laws only require leaving behind 2 per acre following a clearcut. 

So next time you visit a forest, go ahead—count up the snags; note the various stages of log decomposition; observe how light filters through the canopy layers down to the forest floor; wrap your arms around a big tree. Though much of the changes that occur in a forest are slow, you can still appreciate the dynamics of the forest if you take the time to pay attention. 

A big Douglas-fir seen along the trail.

Fish Need Forests

As Bill and I followed the trail in a southwesterly direction, we found ourselves hiking just above the banks of the meandering course of the Sandy River. The Sandy River is about a 56 miles long tributary to the Columbia River and, apart from the Columbia, has the highest productivity of salmonid species in Oregon. Efforts to enhance the Sandy River to ensure it can continue to support salmonid populations are a big part of Bill’s current and past work. 

However, while discussing salmon with Bill, he directed my attention back toward the forest. He pointed out a large down tree near the trail. You see, forests are not only important to terrestrial wildlife.  Fish need forests too. 

Down-tree that Bill pointed out during our hike.

Bill explained, trees in streams and rivers, especially those with roots, provide a place for fish to hide from predators. They also disrupt the flow of water—creating a more dynamic stream channel with resting pools, gravel for fish to spawn in, and habitat for invertebrates.

Historically, the Sandy River had many logjams, as logs naturally recruit in stream beds over time. However, much of the logs in the Sandy River were removed by the Army Corps in the 60s as part of flood control efforts. Since then, returning logjams to the Sandy River has become an important part of stream enhancement work today. 

The Log Father

They call me the “log father,” Bill said—a nickname he acquired due to his persistent hunt for large trees needed for stream restoration. However, creating a log jam is not as simple as finding dislodged trees and dumping them in the river.  It takes a lot of planning, engineering, and equipment to get large logs in place and secure them. It is expensive work too! According to Bill, logjams are like icebergs— they are mostly underground. Human-constructed logjams must be secured in the ground in order to function properly, as well as to prevent them from being washed away in a big storm.

Later, after Bill and I parted ways, I spent some time walking the trails along the Sandy River looking for logjams. I was able to spot the top of several just at the water’s surface.

Logjams in the Sandy River.

Dismal to Happy 

After some time, Bill and I reached a small bridge that went over a very small stream.  This creek used to be named “Dismal Creek,” Bill told me, but now it is called “Happy Creek.” Why? Easy! We were standing next to an old floodplain of the Sandy River that had become disconnected during the age of dams, log removal, and channelization of rivers.

Happy Creek was an attempt to bring water back into the river system by restoring one of its side channels. In order to achieve this goal, a culvert was added on the opposite side of the road to collect runoff and divert it to the floodplain—turning dismal creek into a happy water-filled channel, with even happier consequences. 

What are those happy consequences? Well for starters, floodplains make excellent feeding grounds for fish; they also are a great place for fish to escape turbulent flows and find rest.  In addition, floodplains help reduce river pollution by collecting sediments and removing nutrients. Of course, one of the big reasons floodplains are making a comeback is because they reduce flooding and prevent erosion by dispersing flood energy away from areas we want to protect, like homes and businesses. 

Happy Creek.

How to Restore

Bill and I hiked down to the floodplain to observe it more closely. Seven years ago, Bill was the lead on the “Happy Creek Project,” so he was anxious to see how it was doing. When we got down there, he was thrilled to see the channel they had created had water in it. Though there was no way to know if the Sandy River or Happy Creek was the source of water, he was thrilled to see it was still wet this late in the year. 

The floodplain channel still filled with water in June.

Restoration is still fairly “new science.” Bill discussed that even during the initial phases of the Happy Creek Project, plans were easily dismantled as the team responded to nature. For example, at one point during the project, they found Pacific Lamprey in the restoration site. This was exciting news! But it also required the team to adapt their plans in order to protect the fish.

Failure is part of the gig when it comes to restoration work, but along with it further understanding. “I have learned a lot from each project,” Bill told me. Observing and continuing to monitor projects will only reveal more. Bill said that he hoped to see gravel one day get washed into the floodplain here, creating spawning habitat. Will it? Only time will tell.

Looking Out for Fish

Even though restoration results vary widely, scientists do know a lot about what good fish habitat looks like.  We know what fish need. So if you are visiting a river or stream, Bill suggests looking for several features in order to assess its habitat quality for fish. First, he suggests checking the water temperature. Many fish species in the Pacific Northwest require really cold water to survive and reproduce. If the temperature feels good to you, it is probably too warm for the fish. Second, check the turbidity, or how difficult it is to see the stream bottom. Cloudy water is often the result of sediment pollution and can clog gills or smother fish eggs. Third, look for a variety of substrates in the water. Are there logs and boulders for insects to live on? Is there gravel for spawning? Finally, check for man-made barriers, like culverts that may make travel impossible for migratory fish. 

Of course, if you find any of these features missing, you can also do something about it! For instance, joining your local watershed council is a great way to be involved and learn about restoration work you can participate in locally.

Living with Wildlife 

About halfway through our hike, we looped back up to the road and crossed it to join a trail on the other side.  Just before we made the loop, I asked Bill about what he felt were the important issues or topics in wildlife today. His answer really came down to one major theme—education. Most people still really don’t understand the habits of wildlife. We don’t know how to live with wildlife. 

He explained—when it comes to wolves, for example, we have removed them from the endangered species list because their numbers are up. Yet, their distribution is very limited with packs only established in few places. According to Bill, for animals that have this sort of clumped distribution, delisting just doesn’t make sense

Another example Bill offered was with coyotes. Some people really don’t like coyotes and will kill them on sight. Never mind that coyotes are a minor threat compared to other species, but killing them is counterproductive. As Bill described it, coyotes have an innate reproductive trait that causes them to increase their litter size—from 2-3 up to as many as eight pups—when their numbers are threatened.

Then there is the deer problem. Most people don’t worry about deer populations, and may even feed deer—treating them like wild pets—attracting them into suburban and urban areas. However, according to Bill, deer are the most dangerous wildlife species of all, with more people becoming injured or even dying from deer-related automobile accidents. 

Risk Perception

Perhaps more than any other species, Mountain Lion threat is most misunderstood. Bill told me about a study he was involved in called CAT: scientists, with the assistance of local students radio-collared 25 mountain lions in order to see how much they were interacting with human populations. They found that mountain lions stayed away from people. The only time a mountain lion was tracked near humans during the study was in a case where a farmer was attracting deer, their primary food source.

Yet, people fear cougars because of a few newsworthy incidents. According to Bill, the result of these reports, and associated fears, means more taxpayer dollars being diverted toward tracking down and killing cougars, often without good reason.

If you are personally afraid of mountain lion encounters, Bill recommends avoiding dusk and dawn visits to areas where cougars have been sighted, especially if you plan to bike or run.

Overall, there are better ways to reduce cougar associated risk that doesn’t involve killing the animal.

A Changing Climate

Speaking of changing public perception of risk, climate change remains a risk worth paying attention to throughout the world, including in the Sandy River Watershed.

As Bill and I trekked through some heavy mud on the trail back to where we parked, he made a point to discuss his concerns with climate change. Bill explained—when it comes to climate change we know very little about how it will impact most wildlife species. We know fish will be profoundly impacted, for example, but the extent of the impact is still tenuous.

But, Bill emphasized, that doesn’t mean we shouldn’t do anything about it. In fact, the Sandy River is a cold water refuge for fish traveling in the Columbia River, making it a priority area to protect from climate change.

Bill’s favorite climate change solution is trees. He said that any chance he gets to talk to a climate scientist he asks them about planting trees, and he always gets a positive response.

Thus, under Bill’s direction, about 1.5 million trees and shrub species have been planted in the Sandy River Delta, with future plans to plant more in surrounding areas. Bill focused on using native trees in the plantings, including regionally native trees, such as madrone, oak, and ponderosa pine, chosen as a way to prepare for Oregon’s climate in the coming decades.

Is it just me, or am I sensing a theme?  Trees are not only essential for both terrestrial and aquatic habitat needs right now, but their importance extends much further—as they play a role in improving Earth’s climate future. 

A patch of old growth found along the trail.

Appreciating Wildlife

During the last stretch of the hike, I asked Bill one final question, how can we be more like him? How can someone start thinking and behaving more like a wildlife habitat biologist? Bill had a lot of great answers (some of which I mentioned in earlier sections). However, one idea that stood out as significant was the need to spend quality time appreciating nature. 

Bill emphasized the need to spend time in stillness and silence. He talked about a youth education program he was involved in years ago. One of the activities was a 15-minute silence-solitude station. He talked about an eight-year-old who was determined to remain still during the activity—she didn’t move even when it was clear something was creeping up behind her. Eventually, she was face to face with a doe. “That kind of experience stays with you forever,” Bill emphasized. “One-on-one experiences with nature are invaluable.” 

I tend to agree.

Though I draw the line with mosquitoes. Sorry, Bill! I just can’t!  

Bill Weiler worked for 20 years for the Washington Department of Fish and Wildlife. He now works full time with the Sandy River Watershed Council and as a wildlife habitat consultant.  He is also the author of the book, “Don’t Run From Bears: Living With Wildlife in the Columbia River Gorge.”

Watch your Boots if you Hike with Newts

Watch your step! Rough-skinned newts are on the move this time of year in the valley’s of Western Oregon.  

Rough-skinned newt in Beazell Memorial Forest

Don’t Crush a Newt

A couple of years ago, while hiking with my daughter in Beazell Memorial County Forest, King’s Valley, OR, we discovered a trail littered with rough-skinned newts. Dozens upon dozens of all shapes and sizes, walked clumsily across and past us on the trail.  They were so abundant that we needed to watch our step to avoid crushing them.  Quickly, our ordinary hike in the woods was becoming an unforgettable wildlife adventure. We reveled in the spectacle.  

This year I decided to head back to Beazell to see if the newts were out and about again. And though I was unable to replicate my 2018 experience, the visit got me thinking about the life cycle and circumstance of a rough-skinned newt.  I see rough-skinned newts perhaps more than any other amphibian in Oregon.  Yet, that fateful day in April was something special.  As I trudged up the hill to reach the south meadows of Beazell, I resolved to learn a bit more about these charismatic, orange-bellied creatures, and what sort of mischief they got themselves into. 

Hike at a Glance

  • Trailhead: Beazell Forest Trailhead
  • Distance: 3.9 miles
  • Elevation gain: about 800 feet
  • Notes: There are many options for loops here. You can go a bit longer or shorter depending on your energy. Easy parking and restrooms on site.
View from the south meadow

Death by Newt

Don’t be fooled by rough-skinned newts’ seemingly good-natured demeanor. They may appear benign, but these newts have a seedy underbelly- a very orange seedy underbelly.  Let me explain…

The story goes that in the 1960s three hunters from Oregon were found dead sitting around a campfire with no sign of struggle or injury. The only clue to their death was a coffee pot with a rough-skinned newt curled up inside.  It is thought that the pot, newt and all, had been unwittingly used to prepare their morning coffee- killing the men. 

Toxic Orange 

Though rough-skinned newts are generally a rather drab color of brown on top, they have a bright orange underside. Bright colors are commonly found in the animal kingdom when an animal is trying to make a point- that point being- “I am incredibly toxic so you better leave me alone.” Think, poison dart frog, and you get the picture.  

Rough-skinned newts are no exception.  In fact, rough-skinned newts have a reputation as the most toxic amphibian in the Pacific Northwest; possibly the most toxic on the planet. They produce a neurotoxin called tetrodotoxin (TTX) that blocks voltage-gated sodium channels, important neural pathways. Hence our dead hunters. 

Their toxic orange skin might help explain why newts aren’t particularly evasive too. Just a quick flash of their orange underside (a move called the unken reflex) is an informative gesture meant to deter any predator that might attempt to consume it. Many have tried.  Many have failed. The message is simple- “drink the coffee”- aka eat newt- and you too will face a bitter end. 

Small juvenile newt during migration

The Race

Interestingly, the production of TTX in newt populations has led to an evolutionary arms race with common garter snakes- rough-skinned newts’ only significant predators.  Garter snakes adapt to the poison, but lose some of their prowess. Research has shown a drop in crawl speed in snakes that survive newt skin poisoning. Apparently, the trade off is worth it- evolutionarily speaking.  

However, with more recent research into the source of newt TTX, a third organism has become part of the picture- bacterium.  Though it is difficult to confirm sole responsibility, recent studies have found that some species of bacteria that reside on the skin of a toxic rough-skinned newt are capable of producing TTX.  This suggests the intriguing possibility that our newt is part of a sordid co-evolutionary 3-way yet to be fully understood.

Watery Beginnings

Rough-skinned newts start their lives in water.  Eggs are laid individually and anchored to the underside of leaves or other debris. Upon hatching, rough-skinned newts will spend at least three months as larvae with bushy gills until they metamorphose into adults. Some will never metamorphose- a phenomenon known as neoteny- and simply live out their days in perpetual youth; sort of like a 30-something living in their parent’s basement- why move?

However, most rough-skinned newts will eventually move to a more terrestrial existence. Here they spend much of their time resting under the cover of logs, rocks, or other surface objects, or foraging for food. It is not unusual to see the proud swagger of a newt looking for a tasty invertebrate, especially following a nice rain.  

Plunkett Creek in Beazell Memorial Forest

Springtime Madness

Then, with the onset of spring, and a particularly warm rain- an instinct is triggered in the newts -and it is time to move!  The migration of rough-skinned newts is a heroic spring ritual, as they make their way in droves out from their winter hiding places, toward their breeding grounds. They will travel miles if necessary to make it back to the same pond or body of water to breed year after year; each time following a similar migratory route. 

It is thought that males will generally travel individually, while females have been reported to travel in large groups during migration.  It is possible that the April 2018 my daughter and I experienced was just that sort of event- a gathering of females in anticipation for the “big night.”  To stretch the analogy further, during mating season, males will also exchange their rough, bumpy skin for a more polished appearance- putting on smooth supple skin, a tall tail, and black pads on the soles of their feet. 

Newt during migration on April 28, 2018

The Dance of a Lifetime

When a female arrives at “the dance” (let’s call it), she is swarmed for attention, eventually finding herself locked in a close tango with a single male. This underwater dance can last for several hours.  Then, before the “night” ends the male will drop a package containing sperm (a spermatophore) behind for the female to pick up. If accepting of the gift, she will store it in her reproductive organs. A few days later, when the time is right, the female will then use the stored spermatophores to fertilize her eggs and deposit them one-by-one, preparing a new generation to dance. 

Watch your Hiking Boot

So next time you hit the trail on a warm, wet spring day, keep an eye out and tread lightly. You just might find yourself encircled by a herd of rough-skinned newts. Emboldened by their bright orange belly, they will brazenly follow their chosen route. Not even the crushing force of your hiking boots will hinder them on their path.

Curious at Coyote Wall

View from Coyote Wall of Mount Hood and the Columbia River

A Plethora of Curiosities

One of my favorite hikes in “The Gorge” takes you through a Missoula flood inflicted scablands of oak and pine, up a ridge of columnar basalt, and through fields of wildflowers.  Oh and did I mention, views of Mount Hood and the Columbia River. There is a lot to appreciate along the Coyote Wall trail near Bingen, Washington. So today, let’s explore a few trail curiosities that can be found along the Coyote Wall trail. 

The Hike at a Glance

  • Trailhead: Coyote Wall Trailhead
  • Distance: 7.8 miles
  • Elevation Gain: about 1900 feet
  • Notes: No parking pass required, but popular trail so get here early.  Trail is shared with mountain bikers. Pit toilet at the trailhead.  

The Wall (not just a Pink Floyd Album)

One of the most obvious and interesting curiosities to discover at Coyote Wall is the wall itself.  Formed from ancient lava flows that flooded the area about 16 million years ago, the resulting basalt rocks underwent folding and faulting, and later uplift (both of which continue today), creating this magnificent geological feature. You can see Coyote Wall from the parking lot and again later when you climb up and back down it.  It truly is a wonder and a highlight of this hike. 

If you are so Inclined

You see, Coyote Wall is part of the Bingen Anticline- where the earth’s crust has been compressed, folded and uplifted by faulting. The Columbia River corridor east of Hood River is characterized by convex ridges (anticline) and concave valleys (synclines) formed from a north-south compression of the Earth’s crust. To understand how this works, take a flat piece of paper, or other flexible material, and bring its opposite ends together- the paper will “deform” much like the deformation of the Earth’s crust under similar strain.

As part of the Yakima Fold Belt, the Bingen Anticline is asymmetrical. Thus the Coyote Wall ridge is relatively short (maybe a mile or two), compared to the larger associated syncline valley that the town of Mosier occupies across the river (syncline valleys in the area tend to be 10s of miles).  But don’t let it’s length fool you, uplifted Coyote Wall is a steep climb and descent having been uplifted a couple hundred feet! 

Looking back at Coyote Wall

The Labyrinth (not just an 80’s cult classic)

The Labyrinth

But let’s not get ahead of ourselves- first is the Labyrinth!  Before beginning the steep climb up Coyote Wall, a trail to the east leads you through another fascinating geological feature- a channeled scablands. Curiosity number two!  

Throughout southeast Washington, channeled scablands dominate the landscape. Basically, channeled scablands are areas where parts of the soil and bedrock have been torn up, leaving exposed rocks and deep ravines. How did these scablands form?  What happened here? You might have guessed it- water! And lot’s of it.  

Sculpting with Water

During the last ice age 10,000 to 20,000 years ago, massive floods scoured the landscape. At that time, the Cordilleran ice sheet covered large swaths of North American- but it wasn’t static. This ice sheet would periodically inch its way southward, creating an ice dam along the Clark Fork River in Montana.  The water behind the dam would accumulate into a large lake, the massive Glacial Lake Missoula. At roughly 2,000 feet deep, it held about 500 cubic miles of water. Then, periodically, the ice dam would fail, releasing torrents of water and ice. The flood waters tore through Washington and Oregon eroding much of the landscape and depositing materials as far south as the Willamette Valley. 

With many areas of exposed basalt and butte-and-basin topography, the Labyrinth offers a glimpse into the powerful force of these episodic floods.

Wild about Wildflowers

Desert Parsley – Lomatium

Finally (but not least), are the wildflowers!  I am a huge fan of wildflower hikes- and Coyote wall puts on a gorgeous show starting in the early spring.  Among my favorite of the early bloomers (sometimes seen as early as February) are a diverse group of carrot family plants commonly called Desert parsley.  I don’t know how many species of Desert parsley, or Lomatium, can be found along the Coyote Wall trail. But I saw a couple species on my recent visit, and I am pretty sure there are many more- as there are over 70 known species in the west.  Rumor has it they can be difficult to identify. To be honest, I didn’t even try. 

Better than Carrots

Anyway, besides being beautiful to look at, Lomatium also has an interesting history. The tap root of many Lomatium species was both food and medicine to many Pacific Northwest tribes. For example, the Yakama, who once occupied SE Washington, would use the root of Lomatium, also called biscuitroot or kowsh (yes, there are a lot of names for this stuff), to make small biscuits.  The starchy roots of Lomtium were mashed, shaped, and dried in the sun. Then the biscuits were stored for later use.   

I Think… Probably?

Early reports of Lomatium came from none-other-than Merriweather Lewis and William Clark.  Lewis and Clark called the biscuits derived from the root “chapelel bread” and witnessed its preparation and trade. They also reportedly obtained and consumed some chapelel during their journey.  In addition, Lewis collected and described five Lomatium species for his herbarium. Although it seems he too had difficulty distinguishing between species- using qualifiers in his records such as “I think” or “probably” when attempting to identification.  I’m glad I am not the only one. Though the purple Lomatium pictured below is Columbia desert parsley, Lomatium columbianum… “I think… probably.” 

Lomatium columbianum

Get Curious and Explore

In any event, from huge lava flows to massive floods of water to fields of edible vegetation, there is a lot of science and historical curiosities to explore at Coyote Wall.  Botanically and geologically interesting, it is worth a visit. Stay curious!

Credits/Links

Hike with a Wildlife Biologist

Wild about Wildlife

I love wildlife. Watching a bird on the wing or a deer bounding by makes me feel connected to, and appreciative of, the rich web of life on our planet.  Wildlife encounters can also be a source of inspiration and awe. It can be a humbling experience to stand in the majesty of an animal’s presence. 

However, with so many reports of negative wildlife encounters in recent years, with lives lost (both human and animal), the positive experiences of viewing wildlife are sometimes juxtaposed against a background of fear and uncertainty.  The romanticized idea of wildlife and people living in harmony is exactly that- romanticized. By definition a wild animal IS wild and will behave as such.   

Our Love will Survive

Nancy Taylor on the Calloway Creek Trail

As human populations grow and spread more into wild places, we are encroaching into the homes of our wild neighbors. So what can we do? How can we deal with our current situation?  

As I headed to the trailhead to meet Nancy Taylor, ODFW Wildlife Biologist on Valentine’s Day (nonetheless), for a hike in the McDonald-Dunn forest, these questions remained at the front of my mind. Can people and wildlife ever find love again? 

Perhaps I am a romantic, but I believe the answer to that question is YES!   So grab some chocolate, or your preferred hiking aphrodisiac, and join me on a hike with a wildlife biologist. 

The Hike 

Hike at a Glance

Trailhead: Road 540 Trailhead (Parking area right off I-5, opposite ODFW offices)

Elevation Gain: 200 ft 

Miles: 3 miles

Notes:  Additional parking can be found at the Peavy Arboretum Trailhead.  There are many options for adding mileage to the hike. The hike takes place in the McDonald-Dunn Forests, research forest for Oregon State University.  This is a popular hiking area for locals. A Map of the trail system is available online. 

Blurring the Line

From the trailhead, Nancy and I headed southwest into a Douglas-fir forest chatting about what it is like to work for ODFW as a wildlife biologist.  

Nancy explained that her work entails a lot of public relations and outreach. Though her primary duties are with game animals, she is often dealing with reports of wildlife sightings and alerting the public of these sightings.  It has become a large part of her job over the years.

One of the reasons we chose this hike, in particular, is because the McDonald-Dunn forest has become a mecca for wildlife encounters. The Calloway Creek hike is not in a remote area. In fact, part of the trail abuts a street of homes. Yet there are countless wildlife sighting made here and warning signs posted frequently at the trailheads.

Playing Games with my Heart

Wildlife encounters near cities with lots of green space, is not surprising. Forests provide many important wildlife needs, like food, water, and shelter. 

I asked Nancy how the forest we were hiking through ranked when it comes to wildlife habitat. She said it was “not bad.” With a decent amount of browse, nuts and berries it should support species like deer and wild turkey. 

However, though both of us had seen turkey on the trail in the past, neither of us had seen deer in the area before. Perhaps they found better forage in nearby backyards?

Deer game trail- spotted near the start of the hike.

In any event, it didn’t take long to spot a game trail (most likely from deer) running through the forest.

A game trail is a path created by an animal, like deer, through repeated use. Just like people, animals often follow particular paths through an area while they search for food. Even though you may not see the animal, you can gather signs that that were there. Game trails are easy to spot and can be fun to explore while on a hike, especially when the ground is soft enough to reveal the animal’s tracks.

Who are you?

Looking for tracks on another well established game trail.

On one of the game trails, Nancy and I spotted what looked like cat tracks- probably bobcat, based on their size. There were also tons of canine tracks. But they weren’t wolf or coyote- rather, domestic dog tracks. Which begs the question-Are dog tracks wildlife sign? What exactly is wildlife?

Wildlife is any non-domesticated animal- any bird, mammal, amphibian, or reptile that keeps house in the great outdoors is wildlife.  Dogs, cats, cows, and most horses are NOT wildlife. But they can still be fun to look at.

Endless Love? Setting Limits.

I asked Nancy if there are certain wildlife species that are a priority for ODFW.  She said that game species are the priority. Tracking elk and deer populations is necessary in order to set tag limits and manage game populations.  

ODFW and other wildlife management agencies will conduct spotlight deer surveys where they drive around during certain times of the year counting animals at night when they are most active.  For Elk, helicopter surveys are used for a count. 

You may remember from high school Biology class that populations have a natural carrying capacity.  Basically, wildlife populations are limited by their environment as resources are scarce and predators and other threats are an ever present problem.  In the managed world we live in, carrying capacity has become culturally set- based on human needs and desires, as well as the health and well-being of the population.  When it comes to managing wildlife populations, people are a huge part of the equation.  

Roadkill

Another source of information for wildlife biologists on how wildlife is doing is roadkill.  Sadly, another consequence of people moving through places that animals frequent is that they are far too often hit by cars. An Oregon law that went into effect January of 2019, makes it legal for anyone to salvage the meat of a deer or elk that was accidentally killed in a car collision .  A permit must be filled with ODFW within 24 hours of when the animal is salvaged, and everything must be done according to specific guidelines. One such requirement is that the head and antlers must be turned into ODFW within 5 days of salvage. These heads then become a source of important information on wildlife for ODFW.  Age can be determined by looking at the teeth of the animals. Other health conditions can also be examined.  

Nancy told me she had just dissected an elk head that day to look for signs of chronic wasting disease (CWD).  Caused by a prion- a misfolded protein that causes disease in the brains of animals (kinds of like mad cow), CWD creates holes in the brain tissues of elk, leading to strange behaviors, emaciation, loss of function, and even death.  Though not in Oregon yet, this disease is devastating to elk populations in other areas of the country; and there is potential for spread into other animal populations as well, including humans, though no cases have been reported to date.   

My Habitat is Better than yours

Continuing on our hike, we entered a couple of my favorite spots on this trail- 1) a mossy green riparian area dominated by big leaf maple trees and 2) an Oregon White Oak woodland further up-trail.  As I paused to take in the beauty of each of these familiar spots, Nancy explained how valuable these places are for wildlife.  

Riparian areas are incredibly important to wildlife, especially when you think of the disconnected landscapes that wildlife encounter with human roadways and development.  Riparian areas act as natural corridors for animals to move about the landscape. They provide water and a food source for many organisms as well.  Though they can make up less than 1% of the landscape, riparian areas are used by a lot of different species.

Entering the oak woodlands.

Oak woodlands are great because of their abundance of food.  It is all about the nuts and berries! Wildlife signs were much more obvious and abundant when we crossed into this area.  We saw holes in trees and multiple burrows with acorn shells scattered in their entryways.  

Walk on by

Two invasive species found on the trail- Armenian Blackberry and shiny geranium.

As the trail looped back in the direction we came, we entered what looked like heavily impacted area.  There were more invasive species, like Armenian blackberry, holly, and shining geranium, and the forest had less shrubs that would make good browse for deer.  Invasive species, according to Nancy, are problematic for wildlife species for many reasons, they out-compete native vegetation that may be an important food source for some species, and invasives, like Armenian blackberry, can restrict movement for other species, like deer.  Overall, invasive species impact on wildlife is mixed. What might not be a big deal, and perhaps help one species, can really cause problems for another. 

Show ’em Some Love

As we were getting nearer to the end of the trail, I started thinking more about what might be done to resolve problems between animals and humans.  I asked Nancy, what can people do to help wildlife?  

With the biggest problems facing wildlife being habitat loss, Nancy recommended making this the focus. 

There are many simple things people can do to get involved. Nancy suggested planting native plants that deer and elk can feed on, or putting up bird boxes, for example. Remove invasive species on your property, or get involved in a community invasive removal, or help with a native planting. Also, don’t feed wildlife directly.  Dependency on humans for food is unhealthy for a wild animal. 

You can also support laws and initiatives that put habitat conservation at the forefront of policy.  According to Nancy, in Oregon there is a need for forestry policies that ensure better forest habitat.  Nancy shared her concern regarding the loss of understory plants from plantation cuttings.  Being involved in movements to improve forestry practices is another way to help on a larger scale.

Get to know ’em

More research into understanding population interactions and growth, especially for both our predatory species and game species, can also help.  Understanding how animals move across the landscape through fragmented habitat can inform management decisions.  

Wildlife corridors is another consideration, though an expensive one, for helping wildlife deal with our impingement into their range.  At the same time, putting up fences to help keep animals off our own property can help prevent possible negative wildlife encounters. 

Be Safe

Which brings me to another vital point. If we are going to rekindle our wildlife romance, we need to respect wildlife.  Assume wildlife is around you, even if you don’t see it. Know what to do to be safe.

To put it in perspective, cougar populations are over 6,000 with a statewide range in Oregon.  For safety, avoid being out alone during twilight hours and early morning, especially in areas that have good habitat nearby- and don’t run in these places.  Even in a neighborhood setting, be alert and aware of your surroundings at times when cougars are active. If you do see a cougar, make yourself big. If you are attacked, fight back. And again, don’t run. 

I’ll be Watching you

Just for fun- the octopus tree.

We didn’t see any cougar on the trail (though one may have seen us).  However, not long before we made it back to the trailhead, movement in the trees overhead to our right caught my attention. Several birds flew across the trail, landing on some trees to the left, just ahead of where we stood. It was a varied thrush and brown creeper! This was our first and only wildlife sighting during the hike!  The brown creeper entertained us for a moment by climbing up the tree trunk, while the varied thrush perched nearby showing its burnt-orange markings. Then, as quickly as they arrived, they were gone.  There was no time to take a picture.

When on a trail, watching for movement, like that of our bird friends, is one of the best ways to spot wildlife.  There are a lot of wildlife guides out there for those that want to identify birds or tracks/scat, but just paying attention to one’s surroundings, both sites and sounds, is a great way to start to find and appreciate wildlife. Evaluate the habitat potential for the places you frequent, as well, and be on alert in transition areas like the riparian forest or oak woodlands we saw earlier.  Learn about what you might see and how this may change season to season. And if you really want to see wildlife, be out when wildlife are more active- within a few hours of dawn or sunset. 

For the Love of Oregon

In the Pacific Northwest, there is no shortage of opportunities to view a variety of wildlife.  Oregon has eight major ecoregions with unique flora and fauna due to its variable climate and soils.  More habitat diversity means more wildlife diversity. Nancy recommends Cascade Head as a personal favorite hike for viewing wildlife- she once saw a giant pacific salamander on the trail!  She also suggests heading toward the Cascade mountains for hiking. Canyon Creek Meadows is a beautiful hike near Three Fingered Jack that is recommended. 

Signed, Sealed, Delivered

A pile of acorn shells outside a burrow entry way- anybody home?

So get out there and enjoy some wildlife.  Pay attention to your surroundings and be safe.  Notice the habitat that surrounds you and, if you are so inclined, help protect and restore it. It is a privilege to have natural places near our homes and workplaces (for Nancy, nearly a stone’s throw away from her work), but it is not our space alone.  Show some love for wildlife this Valentine’s day month by giving to wildlife what it needs- a little more space to call home.

Nancy Taylor has 17 years of experience working for ODFW out of the Corvallis, OR Office. She has a B.S. in Biological Sciences from Cornell and a Masters in Coastal Ecology from Louisiana State University. Much of her education and background has focused on  wetland ecology and habitat conservation.