Hike with a Mushroom Citizen Scientist

Tiny unidentified mushrooms growing on a stump near the start of the hike.

Tucked in the duff on the forest floor or growing on the stump of a dead tree—mushrooms captivate the imaginations of many, while others rarely give heed. The beauty and variety of mushrooms, as well as their reputations as delicious edibles, sparks interest for many.

Mushrooms are the fruiting bodies of fungi—a diverse group of organisms, best known for their ability to decompose organic matter. An entire kingdom of life, fungi are mostly unknown (all except a few favorites).

Personally, I have long been fascinated with mushrooms. Their colors, textures, and earthy scents have long drawn my attention. The result— a phone camera full of photos of mushrooms and a lot of fungi-focused questions.

Driven by my fungal ineptitude. I reached out to the Willamette Valley Mushroom Society to see if someone there might be willing to help enlighten me. And to my great fortune, Autumn Anglin, Treasurer for WVMS, answered my call.

Autumn has also fallen under the spell of mushrooms, and by her own admission, fallen hard. Ever since her first encounter with the mushroom hunting world in 2016, Autumn has thrown herself into discovering and understanding them.  Today she leads a mushroom study group and is an active citizen scientist, gathering mushrooms for careful study and genetic analysis.

Autumn suggested we meet at Lewisburg Saddle Trailhead, not far from where I live, where she promised to share with me the scientific side of mushrooms. I emphatically agreed. 

The Hike

  • Trailhead: Lewisburg Saddle Trailhead
  • Distance: approximately 1.5 miles
  • Elevation gain: approximately 250 feet
  • Details: Ample parking at trailhead with overflow parking a bit futher down the road. Pit toilet available. No parking pass required. Roads to trailhead are paved.

A Cold Day in November

It was a chilly day in November.  A freeze set in the night before and a dense white fog shrouded the air like a mystery. The Douglas-fir forest lay still in the mist when I arrived at the trailhead.

Autumn was there ready. Armed with a wicker basket and, as I would soon find out, an unmatched enthusiasm for all things mushroom.

We quickly made a start down the gravel trail in the direction of the new growth trail cutoff.  As we walked along, Autumn told me about her start with WVMS, and how her involvement has grown over the last couple of years. From joining mushroom forays to leading study groups, to her most recent work —a nearly complete at-home genetics lab. 

Stump Stop

It didn’t take long, however, before something caught Autumn’s eye. She motioned me over to a stump on the side of the trail.

“There is a lot of mushrooms on this little log right here,” she noted as we both knelt to get a closer look. She pointed out “a little jelly type” mushroom next to another completely different species, just a few inches away. Larger mushrooms grew from the side of the log.

Autumn pointing to one of the many mushrooms we found growing on a tree stump.

The hunt was on. We started looking for mushrooms everywhere. We looked along the side of the gravel road that sloped upward for black cups, a favorite of Autumn’s. Despite our efforts, we didn’t see any black cups. We did, however, see a few gilled mushrooms, including one with a tan, umbrella-shaped cap that smells like sperm.

Autumn holding the sperm-scented mushroom.

Clavulina Rugosa

Next, we saw a small white coral fungus—Clavulina rugosa or wrinkled coral fungus—growing straight up from the forest floor. C. rugosa is short, only about 3 inches tall, with branches that end in blunt tips. It isn’t the most glamorous mushroom, but what it lacks in showiness, it makes up for in reliability.

Growing near Douglas-fir, often in rings, C. rugosa is a common mycorrhizal fungus in Oregon. As mycorrhizal fungi do, C. rugosa provides nutrients and water from the soil to Douglas-fir trees in exchange for the sugars the trees exude from their roots—a symbiotic relationship essential to a healthy forest.

“Our study group calls it our friend,” Autumn remarked. “We have seen it every time we go out in the winter.”

Clavulina rugosa growing amongst the Douglas-fir needles and moss on the forest floor.

How to Know a Mushroom

At this point, we dropped down onto the new growth trail and headed deeper into the forest where we would see even more Clavulina rugosa among the trees.

The distinct characteristic of the C. rugosa made it easy to tell apart from others, but this is not the case for many mushrooms you encounter in the Pacific Northwest. As Autumn put it when asked how many different species there are in Oregon, “We have a lot!”

Thus, identifying mushrooms is a real challenge.

“My background is I am an artist,” said Autumn. “My observation skills being an artist has helped by a lot.”

To know a mushroom, you must look at the details the way an artist does. Look from all angles. Use your senses! Look, feel, and smell mushrooms. Just don’t taste.

“Run your finger over the stem and gill.,” Autumn suggested. “We can’t tell anything just by seeing a cap.”

The underside of a Russula, showing off the gills and stipe.

Mushroom Anatomy

Another thing helpful when working on mushroom identification is to know something about mushrooms anatomy. Autumn taught me a few terms on our hike together.

“Stems are stipes.” The top layer, or “skin,” is called the pileus layer. “Flesh is everything on the inside, but not the gills.” Gills sit below the cap of some fungi and aid in spore dispersal.

Then there are all the terms used to describe gilled mushrooms—the group most difficult to identify. Words like notched, decurrent, forked, tiered—tumbled out of Autumn’s mouth, as we turned over mushroom after mushroom to look at the gills.

Indistinguishable

However, even with careful study of a mushroom in the field, it can still be impossible to tell with certainty some mushrooms apart just by looking at them.

Autumn explained that even the first level of division in Kingdom Fungi, which splits fungi into two main phyla—Basidiomycete and Ascomycete—requires microscopy to distinguish. Basidiomycetes are identified under the microscope by their balloon-shaped reproductive cells called basidia, with spores that develop outside of the cell on small projections.  While Ascomycete cells look more like “sausages with circles inside them,” Autumn described; the circles being the spores that develop internally in reproductive cells called asci.

Luckily, a lot of microscopy work has already been done, so breaking up fungi into phyla is more a matter of looking it up in a reference book.

But it still takes a lot of work to identify a mushroom, and microscopy is still needed in some cases to tell similar mushrooms apart.

Spore prints—spore deposits on paper—is another helpful tool for identification.

Autumn recommends getting a variety of mushroom reference books to aid in identification. Just make sure they work for the region you are interested in studying. A couple of her favorites are “Mushrooms of the Redwood Coast” by Noah Seigel and Christian Schwarz, and “Mushrooms Demystified,” by David Arora.

Blackfoot Polypore and Xylaria Hypoxylon, Oh My!

It was at about this time that Autumn really became animated, as we started spotting one cool mushroom after another. Eyes bright, a cheerful inflection to her voice, Autumn talked about her mushrooms with the tenderness reserved for old friends and loved ones.

“Blackfoot polypore,” Autumn sang out, reaching down to pick up a branch with a caramel-colored mushroom attached. “They only grow on sticks and dead things,” she went on, turning over the mushroom cap to reveal a porous undersurface and its “black foot,” a dark region where the stipe (or stem) touches the Earth.

The underside of a Blackfoot polypore.

Then, moments later she spotted another.

“This is one of my favorites!” Autumn exclaimed, pointing out a cluster of thin antler-shaped mushrooms, black in color with white tips growing on some rotting wood. “Xylaria hypoxylon.” 

“They are really cool,” Autumn explained. “Feel it!” she suggested, “They are rubbery and tough.” I pulled at the dark strips with my chilled fingers—the wiry body of the mushroom held together firmly.

The white tips are covered with spores, Autumn explained. In the winter, X. hypoxylon, also called “candle-snuff fungus,” produces asexual spores that coat the mushroom.

Xylaria hypoxylon with tips coated in white spores.

Decomposers

“The amazing thing about Xylariaceae,” Autumn continued, is that “they decompose more than any other group of mushrooms.”  Xylaria hypoxylon is one of the thousands of mushrooms in this group.  It is our local decomposing superhero. 

Fungi in general are known for their role as decomposers. When I asked Autumn why she thought fungi were important, decomposition was her simple answer. “Fungi and lichen end up making our soil. They are the foundation of our soil,” Autumn elaborated. From death springs life.

Wood Wide Web

Upon reflection, Autumn added in response to my question regarding fungi importance—”They help communicate.”

Scientific understanding of the “wood wide web” has taken off in recent years, as more research has shown just how important fungi are to a forest. Mycelia extend and spread below the ground creating connections between individual tree root systems. Water and nutrients are provided across this network to trees by the fungi, while trees supply sugar and carbon to the fungi and, at least in some cases, to other trees in the network.

“It’s a symbiotic relationship,” Autumn stated. “It is amazing.”Forests are not just a collection of trees. Rather, they are entire ecosystems—plants, animals, insects, microbes, fungi, etc.—connected by interdependence, as well as competition. 

“We should be thinking of forests as whole living beings,” Autumn remarked. We need to see the forest through the trees.

“And we are part of that ecosystem too.”

Be afraid, Be very afraid

As Autumn turned over one of many brown gilled mushrooms we spotted on the forest floor, she spoke:

“People are afraid of fungi,” she said, “but picking mushrooms is okay.” 

She went on to explain that most of the fungi, the important bits, are underground. When you pick a mushroom, it is analogous to picking an apple off an apple tree. Heck, if done right, you might even help spread its progeny.

“Use a basket or bag with holes to spread spores,” Autumn suggested.

Of course, there are some limitations. “Don’t rake up the forest,” Autumn instructed. But if you find something edible or want to study it, picking a mushroom is acceptable. 

Gear

Autumn carried her own beautiful wicker basket on her arm, while we hiked. 

She stopped to show me the contents: pen, notebook, UV flashlight, whistle, paper lunch bags, wax bags, and a tackle box.

Autumn explained that in order to do a scientific study of a mushroom, she tries to get multiple samples at different ages.  The tackle box allows her to separate out smaller samples, while the wax bags and paper bags are for larger specimens. “Never use plastic bags,” she cautioned. “They will turn to goop.”  

Autumn’s mushroom gear sitting next to a blackfoot polypore.

Poison

Still off-trail, we noticed a tall, white shaggy stemmed mushroom.  The edges of the cap were also jagged, remnants of a veil — a thin membrane that covers some mushrooms before they are fully grown.

Lepiota,” said Autumn. “Poisonous.”

All members of the Lepiota genus are poisonous. Many mushrooms are poisonous, even deadly, another reason people often steer clear.

But, according to Autumn, “all mushrooms are fine to touch.” You just need to wash your hands before eating or touching your mouth, she advised.

“Chemical compounds in fungi need to be metabolized to be dangerous, Autumn explained.

This shaggy species of Lepiota is poisonous.

Russulas

Autumn and I headed to the other side of the tree-lined path. We kicked at a few of what looked like “mush humps”—areas where mushrooms have raised the soil as they begin to come up—but didn’t see anything growing underneath

However, one type (or genus) of mushroom we saw a lot during our ramble, and are often found hidden in a mush hump, are Russulas.

“They are one of the most common, prolific mushrooms,” said Autumn. They are also “under-identified and misidentified,” despite their commonality. Not a lot of DNA analysis has been done on the group. 

The ones we saw had shiny, soft pink caps, but this is not true of all RussulasRussulas come in a variety of bright colors. Autumn picked one to point out the identifying features.  “The main thing is there,” Autumn pointed to the bottom of the mushroom, “gills are attached right to the stem.” This is the signature of a Russula.

Other features include white or yellow stipes (stems) or gills, and a “skin” on the cap that peels back.  Another interesting feature of russulas is their unique flesh—spongy rather than stringy. Their stems “snap like a piece of chalk.” And apparently, if you throw a russula against a tree or similar surface, they “explode.”

One of the many Russela’s seen growing on the forest floor.

Unknown

Autumn and I continued our off-trail adventure in the area of the forest where she discovered a new mushroom just last year.

The fungus itself was first cataloged genetically from a soil sample in the 1970s and 80s but was never described further. Autumn had seen and documented the fungi’s fruiting body for the first time!

As exciting as the discovery was, however, there was still more work to be done. The fungi would need to be found again, Autumn explained, “five or six more times” before it can be named. 

Our eyes scanned the duff, sweeping the area slowly as our feet sunk into the spongy earth.

More to Discover

Surprisingly, very few of the mushrooms on the planet have been identified, so discoveries like Autumn’s are not all that unusual. Experts estimate that as much as 90% of mushrooms out there are waiting to be discovered and given names.

“Most of our mushrooms are European,” Autumn said, “the U.S. doesn’t have that many named species.”

This is problematic because it is likely that the North American version of what we think is a “European mushroom” is probably genetically different from its European counterpart. Work needs to be done to find these differences and, thus, discoveries to be made.

Look Closely

And discover we did! Not anything new to science, but plenty more striking and fascinating mushrooms. 

We saw “stunning” Marasmius plicatulus with their velvety, chestnut orange caps and thin, dark mahogany colored stipes. “They are really sturdy mushrooms,” Autumn noted as she picked one for closer examination, revealing its widely spaced gills. 

The “stunning” Marasmius plicatulus.

Then there was the small, understated Inocybe geophylla mushroom —with its mostly lilac-colored cap — a yellow spot at the center. Inocybe mushrooms have gills that are detached from the stipe.

Silky smooth Inocybe geophylla.

We saw (and I later tasted) a jelly fungus with the common name “cat’s tongue,” a member of the Pseudohydnum genus. The mushroom was translucent with the texture of a gummy. It also had small teeth on the bottom of the cap, giving it a tongue-like look. Autumn offered me a bag to collect some in. I washed them off the best I could before taking them home to prepare and eat. 

The edible “cat’s tongue” mushroom.

Finally, growing on a Douglas-fir cone, a new favorite—the small, but stately earpick fungus, Auriscalpium vulgare. Itty-bitty and brown, with teeth that look like fuzz on the bottom, the earpick fungus immediately stole my heart. How have I never seen these! Autumn told me that they grow exclusively on Douglas-fir cones.  I will never look at a cone the same way again.

Auriscalpium vulgare growing from a douglas-fir cone in the duff.

Mycologist Pace

At this point, Autumn and I had probably hiked about a quarter-mile on the forested trail.

I learned quickly — you can’t really “hike” with a mycologist. It is more like a focused crawl.  You bend over a lot.

We laughed at our slow progress, as we headed back onto the main trail to pick up the pace (only slightly).

For Every Season

Autumn and I followed the trail through the young forest, the mist contrasted against the unpruned trunks of the skinny trees and large stumps. A chill hung in the air.

As we walked, we talked, keeping a watchful eye out for more mushroom gems. Autumn had hoped to find one of her favorite groups of mushrooms, black cups. But today they eluded her.

“It must be too early for black cups,” she remarked.

Too early. Too late. Mushrooms are picky. Some more than others. They pop up when conditions are right, and just as quickly, they disappear.  Then again, they may not come up at all.

Morels are particularly challenging weather watchers.

“Morels need three weeks at 50 degrees before they fruit,” explained Autumn. “If it gets too warm too fast or stays colder longer…” she trailed off, but her face said it all. “Morels are temperamental. They can sit underground for 50 years.”

The misty trail through the “New Growth” forest.

For Every Habitat

We walked past a tree with some cool shelf mushrooms growing on it. I snapped a picture and Autumn grabbed a sample, placing it in one of the compartments of her tackle box.

An unidentified shelf mushroom.

On a dead tree next door, we saw an orange jelly mushroom growing. “Probably not Naematelia aurantia,” said Autumn—a.k.a witches butter.

Witches butter grows on decaying hardwood. This orange jelly was affixed to a conifer. “Dacrymyces chrysospermus—a basidiomycete with two, as opposed to the typical four, spores at the tips of its reproductive cells—doesn’t grow on hardwoods. It grows exclusively on conifers.

An orange jelly mushroom growing on a douglas-fir tree.

Just like other forms of life, fungi have habitat requirements. Knowing where a mushroom will grow can help you tell mushrooms that might otherwise be indistinguishable (or at least without a microscope) from each other.

Planting Biodiversity

We continued down the path. Stopping occasionally to check out a mushroom or two along the way.

Even though we were hiking through a plantation—a forest replanted for timber and not necessarily known for its biodiversity—Autumn noted just how many different types of mushrooms we were seeing.

“Inside plantations, I find that the diversity seems to be a lot more,” she commented.

Biodiversity is important. Biodiversity produces functional ecosystems that are resilient and produce many natural products and provide services humans, among other organisms, rely on.

One of the goals of Autumn and her WVMS Fungi group is to document mushroom biodiversity at each of the study sites.  Using an app called iNaturalist, her group has recorded over 700 observations of mushrooms of a variety of types over a one-year period.

“I feel a sense of urgency due to the climate crisis to get out here,” Autumn went on. She wants to make sure we know what is there!

“It is really important that we preserve this,” she stated empathically.

Chemical Signals

At the end of the new growth section of the trail is a small log bridge that you step through before entering old growth. Here Autumn and I noticed a lot of different mushrooms growing on and around the structure.

“Careful of tan, yellow mushrooms,” warned Autumn pointing to a cluster growing on a decaying stump. “They can be really poisonous.”

Each species of fungi produces its own chemical suite designed to attract, deter, and sometimes kill other organisms. As sessile beings, chemistry is how fungi communicate.  

While the chemistry of mushrooms can be deadly, it can also be a benefit to society. Humans have harnessed compounds extracted from fungi and other sessile organisms as medicines and supplements. Most of the commonly prescribed drugs in the United States, and at least 1/3 of medicines globally, are based on naturally sourced chemicals.

“I never eat white gilled mushrooms,” Autumn confessed. The chemistry is just too risky.

Waxy Caps

Stepping over the bridge, Autumn and I were met with clusters of orange-capped mushrooms.

“Waxy caps,” Autumn exclaimed, “one of the most stunning mushroom genera.”

Waxy caps, or Hygrocybe, are visually stunning. Autumn pointed out how the vibrant cap color contrasted with white, widely spaced gills.

Hygrocybe is also mycorrhizal with Douglas-fir—networking with the roots of trees. However, from where we still the nearest Douglas-fir tree was maybe 100 yards away.

“Does the network go all the way out here?” Autumn wondered out loud.

We were baffled. And impressed. Why not?

A vibrant cluster of waxy caps.

Old Growth Delights

Walking through the old growth forest, was like entering a different world compared to the plantation we had hiked through earlier. The trees were larger and more varied in age and species—including Douglas-fir, Western Hemlock, Big Leaf Maple, and Pacific Yew. Leaf litter covered much of the soil, and shrubs, ferns, and small trees added vertical structure to the scenery.

The trail through the old growth forests.

And then there were the mushrooms!

Bird nest mushrooms with their spores tucked away inside little cups, only to come loose with the splash of rain. A Marasmius mushroom with its bell-shaped cap and fragile stem. And a cluster of brightly colored Hypholoma with black spores and gills that turn green under UV light, as demonstrated by Autumn.

Hypholoma glowing green under Autumn’s UV light.

The List

We continued hiking the muddied path, stopping to look at mushrooms as we discovered them. I couldn’t believe how much we were seeing. A mushroom newbie, I asked Autumn if she had a recommendation for mushrooms to learn first.

“I would say try not to get overwhelmed,” Autumn replied. “You will see a lot of things you won’t be able to identify…try and identify to family and genus to start.”

So, what are the top groups of mushrooms Autumn recommends learning first? Here they are in the order mentioned.

First on the list, Red-banded polypore, or conks. Widespread wood-eating decomposers, red-banded conks are identified by the orange to red stripe that runs between the darker inner and lighter outer edge.

Red-banded polypore found along the trail.

Next, is Mycena—small in stature but large in group size. Mycena is a genus of tiny mushrooms with translucent, fibrous stipes and white gills “you see everywhere,” according to Autumn,

Agaricus is a common urban mushroom. These white to brownish mushrooms with “super tiny, closely spaced gills.”

Stereum is another—a genus of mushrooms that grows on decaying wood. Autumn and I found some growing on hardwood log next to the trail. These thin shelf mushroom clusters are decorated in wavy bands of greens, yellows, and browns. 

Russulas made the list; along with Xylaria hypoxylon and bird nests.

Colorful bands of Stereum growing on a decaying log along the trail

Perfection

As Autumn contemplated what else to include on “the list,” we walked and observed. We saw a cluster of mushrooms (too non-descript to identify), followed by a large single mushroom coated in white (it had been parasitized). 

A parasitized mushroom.

Suddenly, Autumn and I were stopped dead in our tracks. In front of us was a large amanita.

“A perfect specimen,” Autumn spoke, her eyes sparkling. 

I took a few pictures before Autumn pulled the Amanita vaginata, or Grisette, from the earth, revealing its true height.

Amanitas are mycorrhizal species—again, part of the wood wide web.  The mushrooms vary in size from small to large. They have white spores, and their stipes and caps are usually decorated with veil remnants.

As mentioned earlier, veils enclose some mushrooms when they first emerge, later expanding and breaking apart. In Amanita, this means patches or warts on the older caps, and skirts or scales on the stipe.

Amanita can be dully or brightly colored, like the well-known Amanita muscaria, and their caps flatten with age. The Amanita cap has a distinct margin and bell shape with a “nipple” or umbo at its center.

“That is gorgeous!” Autumn exclaimed, overcome with the beauty of it.

We added amanitas to the list. After all, Amanita muscaria is “the poster child of mushrooms,” said Autumn.

The perfect specimen of Amanita vaginata!

Puffballs

After our Amanita discovery, the trail veered back onto a gravel forest road that would take us back to the trailhead.

Autumn’s head swerved around looking for earth star mushrooms along the road—a type of puffball mushroom with a round “puffball” center in the middle of star-shaped rays.

We didn’t find any earth stars, but we did find a few puffballs further down the road hiding in the grass. Though difficult to tell apart. Puffballs are wildly entertaining. Poking a puffball will send out clouds of small brown spores.

Puffballs were also added to the list.

Puffballs found hidden in the grass.

Ascomycetes

We still had some more searching to do as we neared the trailhead.

Autumn was determined to find Helvella vespertina.

“Ascomycetes are my favorite,” Autumn told me, and Helvella vespertina is one of the largest mushrooms in the phylum. (Most ascomycetes are very small and usually lichenized, partnering with algae or cyanobacteria.)

It didn’t take long before we ran across the structural beauty of Helvella vespertina. The first ones we found were parasitized with a white fungus, hiding its black cap, but not its unique shape. Later we found an untouched sample.

It is hard to describe Helvella verspertina—the stipe and cap are wrinkled and folded making it look almost “brainy.” Autumn called them “bizarre and consistent,” an equally apt description.

Morels are the other large ascomycete mushroom with a similar vibe. We added both to our ever-growing mushroom list.

The sculptural Helvella vespertin mushroom.

Mycology on the Mind

Still with several hundred feet to make it back to our car, Autumn and I continued to hunt mushrooms.

We found several more interesting specimens including a Lactarius, or milkcap mushroom that when you break the cap oozes out a white milk-like substance.  

While we hunted, we also chatted more about Autumn’s citizen science work. Thanks to her passion for mushroom biodiversity documentation, her research group earned a grant to send in 50 species this year for DNA analysis to FunDiS, or Fungal Diversity Survey, a nonprofit focused on fungi conservation. 

Now, she plans to continue her work by doing her own DNA analysis of fungi she finds, and by collaborating with researchers and interested groups. With a shoulder shrug, she even hinted at the prospect of going back to school to study fungi.

Mycology Matters

After spending the morning hiking with Autumn, it was clear to me just how much mushrooms and fungi are a part of her life. Her interest and enthusiasm were infectious, and I couldn’t help but smile every time we ran across another mushroom that caught her attention.

I have always enjoyed mushrooms, but Autumn reminded me of just how much fungi matter. It is easy to forget the work fungi are doing in the soil, to take them for granted, but when mushrooms appear they serve as a reminder of the important role fungi have on the planet. 

Seeing mushrooms is like seeing into an otherwise hidden world. So, take the time to get to know a mushroom or two.  Perhaps, learn their names. And give thanks that there are those out there, like Autumn, ensuring their conservation.

Autumn Anglin is a mixed-media artist and vice president of the Willamette Valley Mushroom Society. She is also a “mushroom” citizen scientist, contributing to various mycology studies, including the Fungal Diversity Survey (FunDiS).

Hike with a Beaver Ecologist

Alsea Falls from the lower viewpoint.

Beaver! A surprisingly loaded word. The largest rodent in North America. Oregon’s state animal. The American Beaver is touted for its remarkable ability to engineer waterways. While simultaneously villainized as a nuisance species. Trapped for its fur well into the 19th century, this activity still occurs today, though not to the levels seen during the fur trade.

There are a lot of strong opinions about beaver. They are both beloved and hated. Removed and reintroduced. Marveled at and frowned upon. Yet, for all the attention they get, there is a lot we still don’t know about them.

This is why, after a long day of teaching high schoolers, I met up with Vanessa Petro, who has been studying the American Beaver for over 10 years, to walk and talk about these surprisingly enigmatic, charismatic creatures.

The Hike

  • Trailhead: Alsea Falls Trailhead
  • Distance: 2.4 miles (w/shorter and longer options available)
  • Elevation Gain: 300 ft
  • Details: Ample parking and pit toilet available at trailhead. Drive to trailhead is on well-maintained gravel roads. $2 for parking or use National Forest or other Recreation pass.

The Drive

Vanessa and I drove out to the trailhead together. As we rode along, we chatted about various aspects of our lives—from childhood to career to motherhood.

Vanessa spent her childhood in Pennsylvania, surrounded by nature and the outdoors. “I grew up going out the backdoor and disappearing into the woods,” said Vanessa, much to the chagrin of her parents.

But it wasn’t until she encountered a magazine article featuring Dr. Gary Alt, a renowned black bear biologist, that she initially got hooked on wildlife.  Vanessa saw what Dr. Gary Alt was doing, and as she put it, she “wanted to do something like that.”

Vanessa’s passion for wildlife and the outdoors continued into high school. She was active in hunting during her teen years. And she enjoyed classes in the biological sciences. In particular, Vanessa mentioned an ecology teacher that inspired her and also invited her to compete in a state-wide natural resources competition called Envirothon.

College and Career

Vanessa’s was on a path. After high school, she attended Sterling College, “the smallest accredited college in the United States.” The college only offers environmental study-based majors. Vanessa studied conservation ecology.

“I lucked out while I was there,” Vanessa explained, “There is a large emphasis on hands-on experience. “By the time she graduated from Sterling, Vanessa had already completed several seasons of summer and winter field work, including an internship with Sequoia National Forest working on a forest carnivore monitoring project.

After graduation she bounced around the U.S., living in 11 different states, as she continued to find seasonal wildlife work where she could. “I have no regrets,” Vanessa smiled, “except for maybe a few men.”

Eventually, Vanessa settled at Oregon State University to conduct graduate research.  And once graduated, continued her research work at OSU, which is where she is today. 

Now, married with one child and another on the way, Vanessa spoke adamantly about the challenge of balancing a career with family. “In natural resources, you can’t really get married and have children at the start of your career,” she said pointedly. “I had to wait.” Even today, with a supportive boss and colleagues, Vanessa spoke of the difficulty. “There is a high attrition rate for women in STEM fields,” she noted.

Vanessa posing at the upper Alsea Falls viewpoint.

Why beaver?

Before long we pulled into the parking lot for Alsea Falls and our trailhead. And after a quick restroom break, Vanessa and I headed down the muddy path to Alsea Falls.

Vanessa has been studying Beaver in the Alsea River Watershed for over 10 years. So, as we began our descent, I needed to ask— “Why beaver?”

“For me, I am definitely split between terrestrial and aquatic systems,” Vanessa responded. “And they are in between.” Beaver are semiaquatic—spending part of their time and on land and part of their time in water—taking advantage of resources in both environments.

In addition, many people care about beaver. “There is a lot of conservation interest in beaver in our area,” Vanessa went on. “And there is a lot we don’t know about them.”

Unknowable

Despite their dominating presence in the Pacific Northwest’s history, we don’t even know the most basic information about beaver in this region. “No one can tell you long they live or how many there are in our state,” explained Vanessa.  Their ecology is a mystery.

So, what is happening? Most studies center around using beaver in restoration and/or the relocation of beaver. Questions are very specific and very limiting. For example, “we don’t know the average home range sizes of beaver throughout our state,” but we have movement data for relocated beavers. Yet, beavers that are relocated do not behave the same as naturalized beavers, so it would be like comparing apples to oranges. Thus, there is a gap in knowledge.

Alsea Falls

Vanessa and I continued until we reached the first Alsea falls viewpoint. We stepped out onto the bedrock ledge to get a view of Alsea Falls and snap a few pictures. The ground was slick, but that didn’t stop Vanessa from clambering up closer to the falls for a photo when I asked.

We continued down to the lower viewpoint for another view of the falls from a wider angle. I noticed how the autumn leaves opposite the river were striking against the dark greens and greys of the forest. We looked out at the rushing water.

Dammed if you do, dammed if you don’t

“Do you know what is cool about where we are standing?” Vanessa started, looking upstream, just above the waterfall. “This waterfall is a known barrier for salmonid fish.” But that isn’t the interesting part. Vanessa went on to explain that not too far upstream there is a two-mile stretch of river that beaver have colonized, building dams and creating the perfect habitat for fish.

Beaver are considered ecosystem engineers—they create habitat other species, like salmon, rely on. Birds, amphibians, and invertebrates all take advantage of the engineered wetlands, pools, and other habitats built by beaver. The list of beneficiaries is long.

The site upstream of Alsea Falls is so ideal for salmon that at one point multiple stakeholders wanted to build a fish ladder to provide the fish access to it. The idea eventually lost steam, but just the fact it was considered, says a bit about the quality of fish habitat beaver have the potential to provide.

However, it is important to note that not all beavers make dams. They often don’t need to.  And without dam building the ecological benefit of beaver is non-existent. 

If water around their dens and foraging areas is deep enough to protect them from predators, a beaver won’t build a dam.  Beaver innately rely on these cues to know when to build. Below Alsea Falls dam making isn’t as frequent and “habitat is patchy,” said Vanessa.  The cues aren’t there.

The two-mile stretch of beaver dam habitat above Alsea Falls that everyone desires, well, “that doesn’t occur frequently on the landscape,” said Vanessa.

“Salmonid habitat ends right here,” stated Vanessa, matter-of-factly. But beaver habitat, well that is another story. We followed the trail back upstream with beaver in mind.   

Looking down from Alsea Falls toward a log jam.

Beaver Forage

Back at the top of the trail, we took a sharp left onto a bridge that crosses the Alsea River. I looked down at the flowing reflective waters. Vanessa looked out toward the greenery lining its edges with an eye out for beaver sign.

Beaver are herbivores—they eat plants. More specifically, they snip off smaller diameter branches of trees and shrubs, eat the leafy greens and outer layer of bark where the cambium is.

They are picky eaters though. “They will cut salmonberry, red alder branches, willow, and vine maples… but they won’t touch Pacific ninebark,” said Vanessa. “They will eat lady fern,” but avoid stink currant.

Beaver generally stay close to the water and often sit along its edge while consuming the forage they collected.  “They will forage on average about 30 meters (~100 feet) from the waterline,” stated Vanessa, “but will go further out if they have to.”

“They look for the best of the best,” Vanessa told me later during our hike. Though the Alsea Falls watershed provides several examples of suitable beaver habitat, you won’t find them everywhere. “There are still patches of unpalatable vegetation and undesirable habitat that occur throughout the area.” When it comes to establishing a new home, they look for something equivalent to what they had in the past or better.

“There are no beaver here,” said Vanessa, having fully assessed the area and we continued up the trail.

Looking out over the Alsea River bridge.

Forest Diversions

We headed deeper into the woods, following the muddied trail along a ridge above the creek. The trail weaved through the Douglas-fir trees whose branches caught the late-day sunlight in a bright burst of gold.

Seeing the stately Douglas-fir trees reminded Vanessa of her husband, Andrew Merschel, a forest ecologist. “He just defended his Ph.D.,” she said proudly.

Through his research, Andrew reconstructed the fire history of Douglas-fir forests west of the Cascades, similar to the one we were hiking in.  “The assumed fire return intervals are wrong,” said Vanessa. The actual fire history in this region, she explained, demonstrates a more complex reality than what has been traditionally taught. 

It was fun talking to Vanessa about her husband’s forest research. It was a wonderful diversion. But we were there to talk beaver!  “I want to know more about your research,” I said, as we moved slowly up the path.

Trees filtered the light as we walked through the forest.

Unsuitable

We hiked past another unsuitable spot for beaver—a site heavily forested with desirable alder trees but that was also laced with undesirable stinking currant. No sign of beaver.

Then the trail widened, curving away from the South Fork Alsea River, and began following the bank of Peak Creek.  Here our luck changed.

Much of Vanessa’s research centered around identifying beaver activity, so as we neared an access point on Peak Creek, Vanessa led me down to the water’s edge to look for beaver sign.

Grove of alder. No sign of beaver here.

Beaver Sign

Vanessa explained that many people assume that if they don’t see dams, there are no beaver present. But, as Vanessa made clear early on, not all beavers make dams.  A lot of them don’t need to.  So, to assess beaver activity we need to look for other signs of their presence.

Vanessa climbed down to the water balancing on logs to reach out into the creek where some branches of western redcedar hung over the water. She pulled at the branches, inspecting the tips of each branch until she found what she was looking for.

She directed me to come take a look. The end of the branch she held in her hand was clipped with clear beaver incisor marks. Looking at the branch I imagined a beaver grasping at the branch and snipping it off with its big front teeth, then eating the outer layer of the branch, as Vanessa put it, like eating corn on the cob.

Vanessa noted that clipped branches can be found up and down creeks where beaver reside, but they can be tricky to spot if you don’t know what you are looking for.  People often make the mistake of just looking near the waterline, Vanessa explained, but water levels change all the time, so what is unreachable one day for a beaver might be perfect during high flows.

Feeding stations, a collection of cut limbs along a shoreline or in a protected area, and food rafts, a bunch of clippings floating in the water, are other signs of beaver foraging.  

Peak Creek access point.
Beaver sign! Beaver incisor marks found at the end of a western redcedar branch.

Smelly Stuff

Foraging sign is just one clue or indicator of beaver activity. I asked Vanessa if there is anything else to look out for. “Scat and scent mounds,” she replied.

Scent mounds are territorial markers beavers create out of mud and detritus. They essentially pile up these materials along the shoreline or island in the water and deposit castor, or castoreum, a strong-smelling substance released from specialized glands.

“What does it smell like?” I asked.

“BBQ sauce and vanilla,” Vanessa said. She explained that she often brings castor with her to outreach events and asks people what they think it smells like. BBQ sauce and vanilla are the most common responses. “Castor glands are used to provide ‘natural’ vanilla flavor to ice cream,” Vanessa remarked. Later, a quick google search reveals that beaver butt secretions are used for flavoring in many different food products.

Beaver scat, on the other hand, is less smelly. “Imagine little cylindrical balls of sawdust…sitting in a pool of water,” said Vanessa. It is rare to find beaver scat because of their semi-aquatic nature. “You only see it where they are active,” Vanessa remarked.

We didn’t see any scat or scent mounds during our hike.

Beaver Dens

However, a commonplace to find scat, if you find it at all, is near a beaver den.  A beaver den is essentially the home of the beaver. Beavers use their dens to rest, hide from predators, and raise their young

Beaver den structures in the Pacific Northwest are not usually lodges—dome-shaped structures built with sticks and mud—like are seen on nature shows. Rather, beavers in Oregon, and neighboring states, often dig into the banks of a stream or river—creating a “bank den”—often choosing sites under trees with roots that provide an extra element of structure and protection.  This is a more practical configuration in constrained waterways in Oregon than a lodge. Fluctuating water levels also means that “most colonies will have multiple dens.”

Vanessa and I looked to see if we could find a den under the cedar tree that had been munched on, but there wasn’t any clear opening.

I asked Vanessa about the size of the den, as we prodded around the riverbank. “They are about a foot wide” at the entrance, said Vanessa; and “chambers are actually small.” To find the dens, she usually uses a meter stick to wiggle underneath a bank.  “The tunnel will go into the bank and then will cut up at an angle,” explained Vanessa.

Aging Sign

Having sufficiently checked the area for beaver sign, Vanessa and I decided to continue on the trail.  We were hoping to reach Green Peak Falls, the turnaround point for our hike before dark and we were losing light fast.

As we clambered our way up, Vanessa told me more about her research. “Right now, I am conducting a 5-year beaver dam ecology study in this basin,” she said. “We visit the same sites every year to census of all the dams and beaver activity.”

“We will note all the different types of activity,” Vanessa went on… and “give an age status based on the newest sign identified.” She noted how the clipped cedar branch we had inspected earlier had a lot of “black spotting.” This suggests that the clipping was at least a year old. On the other hand, “it if is clear white,” explained Vanessa, which indicates “they are actively there.”

Chatter

As we walked, I noticed a medium-sized log on the trail that had been “chewed into.”  Beaver? I thought. No. Upon closer inspection, the markings had no incisor imprints. The tree wound was human-inflected, Vanessa assured me.

Having seen similar markings before, I asked Vanessa what was going on.  The markings are called “chatter,” she told me. Beaver will sometimes girdle a tree to wear down their incisors that continue to grow throughout their lives, or possibly to bring down a tree to access more food. In this case, the beaver is doing a lot of “chewing and spitting” and “a huge pile of woodchips” will build up at the base of the tree being cut down.

Vanessa and I “chattered” on.

Should I Stay or Should I go

Before long we reached the stairs that lead to the base of Green Peak Falls. We could hear the rushing waters and paused to finish our conversation before heading down.

“There is another thing with beaver,” Vanessa stated, “just because they occupy a site doesn’t mean they will remain there.”

The duration beaver typically occupy a stream reach and why remains uncertain. “They are all over the map on how long they stay at a site,” Vanessa explained.

The movement of beaver within a watershed is something Vanessa is hoping to find out with her future research. “We have some short-term studies,” she explained, but not enough to really understand what is happening in the bigger picture.

Part of the issue is that beaver dams in Oregon tend to be ephemeral. “By springtime, only 20-45% remain intact from the previous fall,” Vanessa remarked about her study in the Coast Range. And “they only rebuild 7-30% of the time.” More often beaver choose to build a new dam in a new location.

For her study area, “The total number of dams has been consistent at the landscape level over time,” said Vanessa. But much of beaver colonization patterns remains a mystery in response to dam failures.  Will they stay? If they stay, will they rebuild? Or will they go and come back later in the year? There is much still to figure out. 

Vanessa told me about a year when she surveyed her sites, and all the dams failed at one of them, but the beavers remained and didn’t rebuild. But a tributary or two over, another group of beavers constructed 16 new dams.  Why? “I don’t know,” Vanessa responded.

But she hopes to find out!

Green Peak Falls.

Waterfall Mischief

Green Peak Falls was raging when we finally made our way down. The light of day was nearly gone, but I attempted to capture a photo of the falls anyway.

“I like Green Peak Falls better than Alsea Falls,” Vanessa remarked. Then turned and wandered over to the water’s edge. Vanessa was in her element as she balanced on rocks and logs in search of beaver sign. “Sometimes you see chewsticks that came down the waterfall,” she sighed.

After a few minutes of searching to no avail, Vanessa joined me, and we stared up at the cascading falls.

“The pool right above us…” began Vanessa, “I relocated a few beavers to a small tributary above it” She went on to tell me how the male of the pair (the female had died) would leave the release site and come down to the pool just above the falls. “He would hang out for days and then go all the way back up,” Vanessa reminisced. He never headed any further downstream. “Maybe he was too chicken to go around the waterfall,” Vanessa speculated.

I asked Vanessa how beaver typically handled waterfalls. “They do go around them,” she said. “Most of the time they will figure out a way to navigate around them.” However, she mentioned hearing about an unsuccessful beaver relocation where the beavers went over the waterfalls and died. “I have never seen the problem,” she said; “Maybe we have a better batch of beaver…”

Hopes for the Future

Darkness was really setting in now, so we decided to turn around and head back. As we walked, Vanessa talked more about her beaver dam ecology study and her goals for the future.

“We are in season four of data collection,” she said. After year five it will be intensive number crunching and analysis. The goal of the study is to help land and resource managers better understand the realized beaver dam capacity of a watershed and the factors that influence dam longevity at the individual dam and beaver site scales. According to Vanessa, the current popular model used to predict dam capacity tends to overestimate, making it seem like problems exist even when there isn’t one.

“No one knows how to best manage watersheds to promote beaver,” she continued. Our headlamps were now on guiding our way over roots and over puddles as we headed back to our cars. But more recently, Vanessa told me, there is money and interest in solving that problem. After the Labor Day fires, federal, state, and private land managers, in coordination with Vanessa’s lab, discussed the need to implement a landscape-level beaver study in western Oregon.  The study will ultimately include three replicate regions—the Western Cascades, the Coast Range, and Southwest Oregon—and use beaver activity surveys to document their distribution throughout these areas, in addition to other methods like radio-telemetry to track beaver movement and colonization responses to forest disturbances.

Ultimately, through this study, Vanessa hopes that some of the most basic questions about beaver may be answered, like what they need to survive, and how fire and land management may affect them.

Beaver Believer

Vanessa and I continued to talk as we walked in darkness. And before long, we were back at the trailhead and heading home. 

Having spent the evening with Vanessa, I really got a sense for her passion for her work, as well as her ability to be discerning when it comes to beaver science. Many people make the false assumption that if we can just get beaver back everywhere on the landscape, we will be okay. They will fix our problems—from habitat destruction to water conservation to even climate change.

But as amazing as beaver are they can’t fix the damage humans have done to the planet. They aren’t superheroes. Though they probably would look cute in a cape. And just like the rest of the planet, beaver have and will be affected by the dramatic changes in our climate and forests throughout the Pacific Northwest. 

So, as nice as it is to sing beavers’ praises, it misses the mark. To truly appreciate beaver, we need to understand them. That is the first step. And we aren’t there yet. But with the help of people like Vanessa, we might finally learn to walk the beaver walk.

Vanessa Petro is a senior faculty research assistant at Oregon State University. Vanessa earned her B.A. in Conservation Ecology from Sterling College in Vermont and her M.S. in Forest Science from Oregon State University.


Hike with a Field Geologist

View of Broken Top and one of the Green Lakes

I am constantly amazed by the power of water to sculpt the landscape. From glacially carved canyons and deep V-shaped ravines to massive floods capable of eroding and depositing sediment over 100s of miles—water in its various forms has shaped the Earth in profound ways.  The impact of water on the landscape can be seen all around us. If we know where to look.

Luckily for me, I arranged to meet up with Hal Wershow, a geologist and expert on reading the landscape, to help me better see and understand water’s influence in the Pacific Northwest. 

Naturally, we headed to the Cascades to a popular hiking spot in Three Sisters Wilderness called Green Lakes.

Hal in his element, enjoying the views and excellent geology!

The Hike

  • Trailhead: Green Lakes Trailhead
  • Distance: 9 miles round trip to first two Green Lakes
  • Elevation Gain: 1,100 ft
  • Details: This trail is very popular and was heavily trafficked until permits were put in place in 2021. A Central Oregon Cascades Wilderness is required from May to September. The trailhead is easily accessible and there is ample parking. A pit toilet is available at the trailhead.

Opening the Flood Gates

The path hastens along next to a fresh flowing creek lined with conifers and dotted with colorful wildflowers.  A few puffy white clouds floated past us overhead as Hal and I began our hike from the Green Lakes Trailhead.

The ground was level with baseball-sized pieces of pumice and other volcanic rocks scattered between bunches of vegetation. “Fluvial is the term we use for sediment moved by water,” explained Hal. The rounded rock and flat ground are signs that water flooded the area in the past.   

This, of course, begs the question—what happened? Short answer—No Name Lake.

You see, No Name Lake was formed by a glacial moraine, or an accumulation of unconsolidated rock, that was carried in and left behind by a receding glacier from the “Little Ice Age” of the 1800s. Then in the 1960s, an unexplained breach in the moraine occurred resulting in a flood. Perhaps some ice or rock had fallen in the lake generating waves that overtopped the dam causing it to fail.

Interestingly, the source of the flood was reported by Bruce Nolf, a geology professor at COCC at the time—a position Hal now occupies.

The waters from that flood would have washed into the area, Hal explained, carrying sediments and debris all the way across the highway we had just driven in on.

Fall Creek flowing through the flat floodplain at the start of the hike.

Snow Going

It wasn’t long before Hal and I, following the creek-side path, entered a more densely wooded area still blanked in snow. It was early summer and winter snow still lingered in large patches on the trail.

Snow accumulation in the Cascades is incredibly important in the Pacific Northwest. As snow melts it seeps into the ground slowly through pores in rock, becoming part of the groundwater. This water eventually escapes back to the surface through springs that feed our streams and rivers. The lag time between precipitation, snowmelt, and water resurfacing is important helping ensure water supply even in the drier parts of the year.

Snow fields were abundant along the trail.

Spring Forward

Hal told me about a project he is doing with his students at CCOC where they are studying the time water spends underground—also called residence time.  Referred to as the “Spring Monitoring Project,” Hal’s students are locating and gathering samples of water from springs in the Central Cascades near Bend. Then they are sending the samples to a lab for stable isotope dating to determine the residence time of each spring.

Stable isotope dating is used for a lot of applications—to date the age of fossils, archeological artifacts, etc. Elements, like carbon and hydrogen, have a different ratio of their respective isotopes depending on conditions and can change over time. For example, all living things contain a ratio of C-12 to C-14 that is constant, but once an organism dies, C-14 will decay predictably, changing the ratio. This change in ratio allows scientists to determine the age of tissue containing artifacts.

Spring water works in much the same way but uses different isotope tracers to figure out how long water has been underground. The time spent underground varies a lot. Water can remain underground for minutes to thousands of years.

“This research is important, especially in the light of climate change,” Hal explained. With increased drought conditions coupled with increasing demands on water resources, it is important that we understand how much water will be available each water year. Springs with long residence times may be more resilient to climate change.

Rushing Waters

Hal and I continued to crunch over frozen hills of snow, watching out for snow bridges, as we continued to pick our way alongside Fall Creek under a canopy of mountain hemlock and fir.

Eventually, we passed by Fall Creek Falls in just a little over half a mile and took a moment to appreciate the raging white waters as they rushed down a short rockface. Fall Creek and its falls are fed by the same waters that fill Green Lakes which, in turn, are fed by glacial and snowmelt from South Sister.

Waterfalls are another example of the force of water on the landscape. Water is an agent of erosion, but not all materials erode equally. For example, most sedimentary rock erodes easily, while others, like igneous rock, granite, are more resistant. Waterfalls, like Fall Creek Falls, form when there is a difference between the materials that make up the streambed. Essentially, the material below the waterfall eroded more easily than the material above it.

We continued to trace Fall Creek’s flow further upstream, the trail trending uphill through some switchbacks, eventually crossing the creek on a narrow log bridge.

Fall Creek Falls as seen from the trail.

Walk on, Rock On

As we walked along the path, Hal pointed out some of the different rocks found along the trail. All the rocks we saw were igneous rocks—formed from cooled magma.  

In general, igneous rocks can be divided into two major groups based on their silica content—mafic rock and felsic rock. Mafic rock is low in silica (45-55% silica) and is generally darker in color. The lava is less viscous (due to its low silica content) and erupts smoothly, as gases readily escape and don’t build up generating the pressure needed for an explosive eruption. Dark grey basalt is a classic example of mafic rock. 

Felsic rock on the other hand is high in silica (65% or higher silica) and tends to be lighter in color. The lava is much more viscous and stickier making it difficult for water and gases to escape. The result is a buildup of pressure and more explosive, violent eruptions. Pale tan or pink rhyolite is a classic example of felsic rock.

Light grey andesite is an intermediary (55-65% silica) between mafic and felsic. Andesite rock has enough silica to produce quartz crystals, so it often has a “salt and pepper” appearance.

Disorganized

However, the chemical composition of igneous rocks is not the only thing that determines their final structure. For example, rocks exposed to oxygen may become redder; rocks that form under the Earth’s surface grow larger crystals; and rocks formed during explosive eruptions may be more fragmented.

One of the most common rocks Hal pointed out on the trail was pumice. Chemically, pumice is like any other rhyolite rock, but because of the conditions it formed in, pumice has some unique qualities. 

Pumice is formed during violent eruptions of very viscous rhyolite lava that is very high in water and gases. When ejected, the gases escape rapidly and the water evaporates and expands, causing the lava to become frothy. Pumice is a disorganized rock—formed so quickly that there was no time for it to crystalize. Hal called it “volcanic glass.”

The resulting rock is an incredibly light, vesicular rock with the reputation of being able to float in water.

Slow your Flow

However, one of the most striking rocks seen on the trail isn’t pumice, but obsidian—a shiny, (usually) black rock, generally known for its use in arrowheads and other edged tools. The cutting edge of an obsidian tool is sharper than a surgeon’s steel scalpel. 

Not too long after crossing Fall Creek, part of the 2,000-year-old Newberry lava flow comes into view—a massive wall of rhyolite—much of it in the form of obsidian. The wall is a spectacular feature for the next few miles, hemming in Fall Creek on the opposite bank from the trail.

The wall of rhyolite starting to come into view.

Hal explained that obsidian, like pumice, is also rhyolite. However, unlike pumice, obsidian is not the result of explosive eruptions, but rather viscous lava that exudes slowly from volcanic vents. Just like pumice and volcanic ash, obsidian has no crystalline structure and is also “volcanic glass.”

Hal described the lava flow as being so slow that the movement would have been imperceptible to the human eye—we are talking less than a few meters per hour.  The flow would have also been cooler and not like the red-hot magma seen erupting from volcanos in Hawaii that tend to be mafic lava flows.

More views of the rhyolite lave flow. The dark, shiny rocks are obsidian.

Lakes O’ Plenty

Hal and I continued to hike uphill through the forest, crossing several smaller creeks as we went. Eventually, we reached a sign with a map indicating we were about to enter the Green Lakes Basin. 

Early in the hike, Hal told me that there were several ways lakes can form. A glacial moraine is one way, like the one that formed Broken Top’s No Name Lake. A lava flow dam is another. Green Lakes is an example of a lava-dammed lake. From the map you could see where the lava flow displaced the creek and cut off most of the area above, creating the basin. 

Hal also pointed out the areas where water is flowing into Green Lakes. Not just water, but sediment too. They are being filled up, Hal explained. The addition of sediment means that Green Lakes will not be around forever.

“Another 1,000 years and they won’t be here,” Hal stated emphatically.  

Stopping to check out the Green Lakes map and sign.

Composite

Past the sign, the first of the Green Lakes comes into view. Flanking the blue-green waters are two massive peaks—South Sister and Broken Top.  Like sentinels, they tower above Hal and me. While at the same time, seemingly close enough to touch.

Both South Sister and Broken Top are stratovolcanoes, also called composite volcanoes—named for the varying nature of erupted materials that build their steep cones—anything from lava to ash. The formation of a stratovolcano is a process of building up and tearing down. They are known for violent eruptions where large amounts of their mass may be ejected into the air—sometimes leaving a large crater. Mount St. Helen’s is a composite volcano. Mt. Mazama, where Crater Lake now stands, is also a composite volcano that blew its top over 7,000 years ago.

South Sister, a relatively young composite volcano.

Fire and Ice

However, as Hal reminded me, volcanism is not the only powerful force at work in the High Cascades. Ice—in the form of glaciers—is also a powerful agent of change in this volcanic landscape.

South Sister, with her tall dome shape retained, is still active—with recent eruptions dating back only a couple thousand years. In contrast, Broken Top is a long-extinct volcano—last active over 150,000 years ago. Since then, Broken Top has been roughly hewn by glaciers leaving its summit a jagged pile of rock and eruption crater exposed. Glaciers are moving ice, capable of abrading and polishing down rock, creating steep-sided hollows, and leaving behind sharp peaks and ridges. Hal pointed out some of the features formed by glaciers on Broken Top, including a cirque, horn, and arete.

Glaciers can still be seen on both Broken Top and South Sister—though they are much smaller and fewer than just a hundred years ago due to anthropogenic climate change. Staring up at South Sister, I asked Hal how to identify a glacier well enough to tell it apart from snowpack. “Crevasses—deep breaks in the ice formed as different parts of a glacier travel at different speeds—are one key difference,” Hal responded.

But Hal also noted that Glaciers can be very difficult to spot. So difficult, in fact, that only a month earlier, a “new” glacier was discovered on South Sister by Oregon Glacier Institute, an organization with the goal of identifying and monitoring Oregon’s glaciers. And by “new,” I mean new to science. “Glaciers tend to be in areas that aren’t very visible,” Hal warned, “making them difficult to locate.” 

Heavily eroded Broken Top

Alluvial Fans

Continuing our hike, Hal and I followed a trail that put us closer to the base of South Sister. Here we reached a deep water crossing and a view of one of the alluvial fans that South Sister’s meltwaters created stretching out in front of us.

An alluvial fan forms when terrain suddenly becomes less steep, like at the base of a mountain, and the water flow less restricted. As the gradient is lowered, the water flow slows and spreads out, dropping sediment in a fan or cone shape.

Earlier in the hike, Hal pointed out a “mini-version” of an alluvial fan where steep flowing drainage of water slowed near the trail as the path of the water flattened and the water was unconstrained. Though perhaps not as dramatic as the large alluvial fan in front of us, the principals are the same. When water slows, sediment drops out.

Hal and I considered crossing the creek to get a better look at the first fan, Hal even attempting to balance his way across some unstable logs, but instead opted for an adventure around the second Green Lake and past the third to the alluvial fan on the far side of Green Lakes.

Hal with a mini-alluvial fan on the trail.
The first water crossing looking out toward an alluvial fan

Round We Go

As Hal and I made our way around the largest of the Green Lakes, we kept a lookout for more geological treasures.

The snow continued to be a bit challenging at times, but we treaded carefully along the narrow trail. 

Before long we spotted signs of an ephemeral spring. Though no water was rushing forth from the Earth, Hal pointed out the eroded channels, changes in vegetation, and exposed roots—all indicators that water had flown forth at some point during the year.

Hal pointing out signs of an ephemeral spring

A bit later, Hal spotted a perfect example of high silica, rhyolite, and low silica, basalt sitting side by side on the trail.

Rhyolite to the left with basalt to the right.

Breach

Eventually, we made it to the bottom of the alluvial fan. Hal explained that there was evidence, at least in part, that the fan was a result of a breach in a moraine-dammed lake further up the mountain. The plan was to head off-trail and follow the alluvium up to see if we could reach the moraine lake.

Almost immediately after heading off-trail, Hal started pointing out the changes in terrain. Like a kid-in-a-candy-store he had me looking at the rock that now littered the ground.  “No pumice!” he exclaimed.

Instead of pumice, the area was filled with volcanic rock that looked speckled—with larger crystals embedded in a finer grain. A “porphyritic texture,” stated Hal—formed from lava that cooled slowly below the surface before rapidly cooling above the surface.

The “fresh rock” signaled to Hal that the lava bed we were walking in was from a different eruption than the pumice and lava flow from earlier.

Fresh volcanic rock!

Signs of a Flood

Hal’s excitement continued as we picked our way up the drainage—the area was literally awash in signs of past flooding.

For one, the size of the rocks changed—smaller rocks gave way to larger rocks—as we moved up. Hal explained that this was expected, as smaller rocks can be carried by the floodwaters farther than larger rocks, which would have been dropped closer to the breach.

Larger rocks also piled up along the edges of the now-empty flood channel—forming natural levees. Again, Hal explained how the energy of the floodwaters would have dissipated toward the edges, dropping these boulders into place.

Hal also noted how the forest looked different in the flood zone. Looking beyond, you could see a lot of taller trees, but within the flood zone, there were only small trees. Trees in the area would have been toppled by the floodwaters. The smaller trees, Hal explained, would have sprouted after the last big flood.

Natural rock levees at the start of our off-trail climb.

Flow Banding and Glacial Polish

Hal and I continued to pick our way over larger and larger rocks. Along the way, we saw some more fun geological features in the rock.

One such feature was a large rock near the edge of our flood channel that looked striped or banded. Hal explained that each band was really the result of different flow rates in the lava that cooled to form the rock—a phenomenon known as flow banding. Flow banding occurs because there is the shearing force between the layers of lava causing them to flow differently relative to one another. 

Hal’s geologist mini-figure sitting atop a flow banded rock.

A bit later, Hal pointed out another rock.  This one was smooth with some well-defined grooves. Unlike the flow-banded rock, the lines in this rock were formed from a glacier. When glaciers pass over rock, Hal explained, they carry gritty sediments that will abrade the rock, polishing the rock smooth.  If a larger rock is stuck in the glacier, it will carve deeper grooves in the rock as well.  The overall effect is called glacial polish. Hal suggested thinking of it like sandpaper—different parts of the glacier may have a different grit resulting in differences in the polish.

Hal pointing out the glacial polish on one of the many boulders along the trail.

Survivor

We continued heading up the rocky drainage, crossing several snowfields. The rock levees are now as much as 10 feet tall in places. Looking back, beautiful views of the Green Lakes Basin periodically caught my attention. 

Apart from the snow, boulders made up most of the ground surface as we trekked upward. The young forest seen toward the base of the washout was nonexistent.  But what we did find were remnants of a vegetative past.

At one point, Hal and I saw a log stuck in the sediment that sparked some interest. Organic material, like the log, can be dated using either radiocarbon dating or dendrochronology. Radiocarbon dating would provide the apparent age of the tree, a decent estimate of age as far as geological events go.

Hal recording video of a log stuck in the sediment.

However, one of my favorite spots on our hike was where we passed a live tree that had somehow survived the floods. Though a bit disheveled, broken and stripped of bark on one side, it was beautiful in its own way. We stopped for a while by this tree, breaking for water. Standing there looking up at its worn trunk I was drawn to its ruggedness. It’s history. It’s story.

A Story

Hal and I never made it to the glacial lake to see the breach. Logistics didn’t allow for it. We did, however, see its effects.

The story of the Earth is one of constant change—often slow but punctuated by quick, sometimes devastating, alternations. Hiking with Hal reminded me of this.

Powerful natural forces that shape the planet, like water, make change inevitable, but also knowable. The story of our planet unfolds as we read the geology. And, like a tree battered by floodwaters, it is one of beauty and resilience.

The survivor!

Hal Wershow is an Assistant Professor of Geology at Central Oregon Community College. His prior experience includes work in the environmental services industry and geoscience education. Hal earned a Master’s in Geology from Western Washington University.

Top Fall Hikes in Oregon for the Curious

Autumn is one of the best times of the year to explore the Pacific Northwest. The air turns crisp and cool, mosquito season is over (a big win in my book!), mushroom hunting season ramps up, and the deciduous leaves of our native shrubs and trees brighten as they change colors. It is absolutely a favorite time of year for me to hit the trail!  But where to go?

Though it is difficult to go wrong when selecting an autumn hike in Oregon, there are several hikes that I have found especially welcoming this time of year. Here are a few of my top picks for my home state to be curious about .

1) Silver Fall State Park

Silver Falls is a classic Oregon hike. It is beautiful any time of year, but fall is one of my favorite times to visit, especially on a drizzly day. Giant bigleaf maple trees turn a bright yellow in the fall and litter the ground along with a myriad of other deciduous leaves of various colors and shapes. When the wind picks up, leaves fly down from overhead before finding their way onto the soft earth or dancing along in the waters of Silver Creek. The smell of the Earth is richer in the rain and who doesn’t love a bit of mud or a good splash-able puddle?

Location or Nearest Town: Near Silverton, OR. Trail of Ten Falls starts at South Falls; parking permit is required.

Distance: 7.2 Miles for the Trail of Ten Falls (highly recommend). There are many shorter options as well. Check out the Oregon State Park’s Silver Falls Map to learn more.

Difficulty: Moderate.

When to go: Mid to Late October is typically best for fall colors. I also recommend visiting when there is a bit of rain in the forecast.

Why go? Fall colors juxtaposed against dark green conifers. Silver Falls is also a great place to look for a diverse array of mushrooms. Oh, and then there are the waterfalls!

Trail Curiosity: Fall Leaves!

Leaves in the spring and summer are rich in chlorophyll, an important mixture of compounds that, not only make leaves green but is essential for plants to perform photosynthesis—the light-requiring process they use to produce food. During the fall, with less light available, some plants will halt photosynthesis and, instead, break down their chlorophyll in preparation for winter dormancy—a process known as senescence. During senescence, nutrients are redistributed away from leaves to seeds and buds for the following spring.

Without chlorophyll around, other pigments in leaves become visible. Classes of compounds, like carotenoids and flavonoids, which give leaves an orange or yellow hue respectively, are unmasked in the fall, while others, like red inducing anthocyanins, may even ramp up production in years with a lot of sun. And voila! We have a forest of eye candy!

Because the timing of foliage color change is triggered by changes in day length and decreased temperatures, the timing of color change is predictable. A tourism site called Smoky Mountains publishes a fall foliage prediction map each year to help tourists get the timing right. Other factors, like, sunny days, nighttime temps, as well as moisture conditions can also affect the vibrancy of the fall show.

2) Ramona Falls

The hike to Ramona Falls near Mount Hood is a Portland favorite. The trail features a mossy green forest and mountain views, culminating with access to the lovely Ramona Falls—her water tendrils cascading down a basalt cliff at trails end. Though the hike is accessible much of the year, except for winter, fall is an especially enjoyable time to visit.  Why? Because relentlessly pushing up from the duff are countless mushrooms of spectacular variety. Tall or short, symmetrical or irregular, purple or brown—there is much to admire about the biodiversity of mushrooms in the area. In short, fall is mushroom season.

Location or Nearest Town: Mount Hood Wilderness area. Near Rhododendron, OR. Start at the Ramona Falls Trailhead.

Distance: 7.1 miles with additional miles possible.

Difficulty: Moderate (Caution: do not cross the Sandy River when water runs at dangerous levels).

When to go: October to early November.

Why go? Mountain views, a crystal-clear creek, and, of course, Ramona Falls.  And in the fall so many mushrooms!   

Trail Curiosity: Mushrooms!

Fungi are ubiquitous. Especially in a forest. Creating massive networks of mycelium, growing on and inside plants and their tissues, floating in the air as spores—fungi are all around us. They are found in nearly every ecosystem on the planet, and we are more closely related to kingdom fungi than kingdom plants. Despite this, we hardly give notice. That is until they mushroom.

Mushrooms are the fruiting bodies of fungi—releasing spores that are carried by wind, or by other means, to a final resting place. If conditions are suitable, a new fungus will begin to grow. Fungi hedge their bets by producing a lot of spores—sometimes numbering in the trillions.

The spores of a fungi can also be used to help identify a mushroom. Many people create spore prints for this purpose. Others make prints as a form of art. Making a spore print involves placing a mushroom cap gill side down on paper, usually half black and half white, and covering it with a glass cup for 24 hours. When time is up and the cap is removed, a ghostly image of the top of the mushroom cap is left behind.   

Besides performing an essential function for the fungi, mushrooms are magnificent to behold. From behemoth king boletes to the dainty (and illegal) liberty caps, the biodiversity of mushrooms in Oregon’s forests is spectacular. There are the popular edible mushrooms, like honey-colored Chanterelles and the fleshy white or brown Matsutake to forage for.  Poisonous mushrooms, like the bright-white destroying angel and the greenish and white gilled death caps, to avoid. And many more mushrooms to admire.

3) Old Salmon River Trail

The Old Salmon River Trail is a blissfully easy river walk. With a gentle grade it is appropriate for all ages and most ability levels. The trail runs along the Salmon River on one side and the hardly used Salmon River Road on the other. The path is flanked by mossy, green vegetation, and towering old growth trees that gives a sense of being more remote than it is. The rush of the river near the trail adds to the tranquility, as does the wildlife that call the watery corridor home. In the fall, Chinook Salmon may be seen spawning in the rushing waters of the wild and scenic Salmon River.

Location or Nearest Town: Near Welches, OR. Start at the Old Salmon River Trailhead.

Distance: 5 miles round-trip or arrange a shuttle for a shorter walk.

Difficulty: Easy.

When to go: September to October to see Chinook Salmon spawn in the river.   

Why go? Old growth trees, mushrooms, and fall color. This hike really has it all, including the chance to see Chinook Salmon spawning in the fall.  

Trail Curiosity: Spawning Salmon!

Chinook Salmon, like other salmonid species, are anadromous fish—born in freshwater, but spending most of their lives in the Ocean, before returning to their natal stream to spawn.  They are also called “king salmon” due to their massive size—sometimes reaching up to 50 pounds. With spotted purple backs and silver sides, they are distinguishable from other salmon by a darkened gum line.

The migration of a Chinook to and from the Ocean is full of challenges—from predators to pollution to dams—very few make it back.  Those that do return, bodies darkened with a deep red underbelly and fins—are ready to spawn.

The female prepares a redd, or “nest”, by flipping her body sideways and slapping her tail against pebble sized rocks. Her silvery sides shimmering, as she cleans out a depression in the gravel.  She lays thousands of eggs into the redd while, simultaneously, a male release his milt—a sperm containing fluid—fertilizing the eggs. Their purpose complete, both parents will die within a couple weeks of spawning. Several weeks later the eggs will hatch, and the cycle will begin again.   

4) Scott Mountain Loop

The Scott Mountain Loop begins at scenic Scott Lake. It is worth walking down to the lake before beginning the loop to view the Three Sisters above Scott Lake’s glassy waters. Once on the trail there is much to appreciate, including a subalpine forest of fir and mountain hemlock, several large swimmable lakes, and excellent views of the surrounding mountains.

In the fall, patches of huckleberry shrubs turn a vibrant red along the trail, contrasting against the greens and yellows of the forest. The trail starts at over 5,000 feet of elevation, so as you move up trail to even higher altitudes be prepared to find early snow lingering on the trail.  Upon reaching the open meadow summit of Scott Mountain, you will encounter views of many of the nearby Cascade volcanoes, including Mt. Jefferson, Three Fingered Jack, Mt. Washington, and the Three Sisters—with any luck, freshly capped in white.

Location or Nearest Town: Off the McKenzie Highway; Near Sisters, OR. Start at the Benson/Tenas Trailhead.

Distance: 7.7 miles for an out and back to Mount Scott summit or 9.7 miles for the loop. A shorter easier hike to Benson Lake or Tenas Lakes is also possible.  Please note you must have a Central Oregon Cascades Permit to hike here.

Difficulty: Difficult with Easy shorter options.

When to go: Late September to Early October after the first snow in the mountains.    

Why go? Swimmable lakes, green subalpine forest, and mountain views!

Trail Curiosity: Weather Transitions!

Autumn is a time of transition—it is fleeting, sometimes bittersweet. The heat of summer wanes in favor of cool temperatures, clouds, and even some snow showers in the mountains. For many it is the best time of year to be outdoors in the Pacific Northwest, claiming that “the weather is perfect!”

One of the major factors that influences climate and weather patterns on the planet is sunlight. The Earth orbits around the sun at a tilt, such that at different times of year, depending on where you live, you may be facing the sun or turned away from it. The Fall Equinox, usually around the 22nd of September, marks the day where the Northern Hemisphere starts to tilt away from the sun as it gradually changes in its position in orbit. Without the direct sunrays of summer, the days grow shorter, and the temperatures start to drop—sometimes rather sharply. This leads to the first freeze and mountain snowfall. Winter is on the way, especially in the mountains.

Hike with a Dune Ecologist

View from one of the dunes in Pacific City, Oregon

There is something otherworldly about walking among the dunes on Oregon’s Coast. Walking from crest to crest of these rolling hills of sand feels akin to walking atop ocean waves. With each step the sand shifts underfoot, and you wonder if you just might comfortably fall into a deep crystalline sea.

Despite the strangeness of the sand dune landscape, dune ecosystems are common along Oregon’s Coastline (though not so much elsewhere). Formed following the last Ice Age by the erosion of the sedimentary rocks of Oregon’s coastal mountain range, Oregon’s dunes are sculpted and shaped by the seasonal influences of water and wind and the surrounding terrain. The result is a menagerie of dunes of varied shape, size, and expansiveness; some only a few hundred feet long, hedged in by headlands and other obstructions, while others stretch 10s-of-miles along the coast creating places like the Oregon Dunes National Recreation Area near Florence.

Given the ubiquitous and curious nature of dunes, it is no surprise that these large sandy masses, have been, and continue to be, influenced by people.

When European settlers flocked to the Oregon coast to establish communities in the late 1800s, they were immediately met with the challenge of living in a dynamic sand-swept environment.  Rather than retreating, settlers managed the sand with what they considered the best weapon in their arsenal—European beachgrass. European beachgrass has been used extensively throughout the world for erosion control. Oregon is no exception. European beachgrass is now a defining character of Oregon’s Coastal Dunes, influencing not only dune formation but also dune ecology.

To better understand the story of dunes in the Pacific Northwest and the ecological significance of European beachgrass and another non-native grass, American beachgrass, I met up with Rebecca Mostow, dune ecologist, at Bob Straub State Park near Pacific City for an interview and hike.

The Hike

  • Trailhead: Bob Straub State Park Trailhead
  • Distance: approximately 1 mile
  • Elevation: minimal
  • Details: Plenty of paved parking at trailhead. No fee for parking. There is also bathroom with flushing toilets. There are signs and a map posted at the trailhead.

It was a cool spring day when I met up with Rebecca for our hike and interview. A marine layer of clouds hung low in the sky, threatening drizzle, but the weather remained mild and dry. After some brief introductions, Rebecca and I headed out along the back dune trail, or Marsh Trail, at Bob Straub Park with plans to eventually cut down to the beach. 

Rebecca Mostow standing on the crest of one of the vegetative dunes along the trail.

Rich and Varied

As we made our way down an alley of shrubs and trees that line the entire back dune trail, Rebecca told me a bit about her background.

Rebecca’s undergraduate work was at Oberlin College, where she studied biology and plant systematics.  She also got involved in a variety of different research projects, from an HIV vaccine research internship to an invasive snail survey.  Her work continued to be rich and varied after graduation as well. She even spent some time on a tiny island with half a million sea birds and only two people for company.

However, it didn’t take long before Rebecca’s passion for plants drew her back. Even during her time watching sea birds, she admitted that she couldn’t help but notice the plants.  “I made a baby flora of the island,” she remarked. She also worked for the BLM in Nevada, Carson City district, on plant conservation before ultimately ended up in graduate school at Oregon State University in Sally Hacker’s Lab.

Intentional Introductions

Even now, walking along the trail, Rebecca was drawn to the plants surrounding us. She admired the new shoots of the spruce trees that grew along the trail, impressed by their symmetry. “See the little pattern,” Rebecca exclaimed, “a Fibonacci spiral!”

As we continued along, however, another pattern became apparent, at least to Rebecca’s trained eyes. “All of the plants we are looking at were intentionally planted altogether,” she explained.  European beachgrass was planted to help stabilize and build dunes, Shore Pine to provide native habitat, and invasive Scotch Broom as a nitrogen fixer.

“They were trying to engineer a coastal forest,” offered Rebecca, as an explanation.

Shore Pine, Scotch Broom, and European beachgrass growing together along the trail.

Unstable

As mentioned previously, sand does not make the development of permanent human settlement easy, or in some cases possible. By vegetating the dynamic, shifting sand system that was once present in the area, it was converted into a more stable system. But at a cost.

Rebecca explained that many species rely on open sand for habitat. Species, like the Western Snowy Plover, require open sand for nesting. The Streaked Horned Lark is another species reliant on open sandy areas.

So, though sand stability is great for human habitation, it does result in a substantial loss in habitat for other species. Both the Western Snowy Plover and the Streaked Horned Lark are threatened species under the Endangered Species Act mainly due to habitat loss.

This Grass is not like the Others

I have never been very good at identifying grasses and grass-like plants. Beyond remembering the saying: “sedges have edges, reeds are round, and grasses bend their knees to the ground,” I have very little experience with grasses and couldn’t begin to tell species apart.

However, walking with Rebecca along the dune trail, literally lined with beach grass, it wasn’t long before I received an education.  

“This is our native dune grass,” Rebecca chimed, pointing to a blue-twinged blade. She went on to explain that in addition to the color of the blade, there are many other characteristics that help distinguish American Dune grass, Leymus mollis, from the European beachgrass, Ammophila arenaria, that was brought in to build and stabilize dunes.

“The leaf blades are so much wider, and it has a more prominent midrib” than European beachgrass, said Rebecca.

Rebecca examining a patch of beach grass.

The Pits

And then there are the pits. Okay, so they aren’t exactly “pits,” but if you look at the point where the leaf and stem meet, you can see something called a ligule, a thin translucent-white tissue growth found at this junction. “The ligule is the armpit hair of the grass,” said Rebecca partly in jest, as she bent back a leaf. But it is also one of the most surefire ways to differentiate between beach grasses. American dune grass has a ligule that is short and flat. European beachgrass has a ligule that is much longer. When you need to tell grasses apart, “Just look at the pit hair!” Rebecca exclaimed.

Growing Underground

After my ligule tutorial, I asked Rebecca why European settlers on the coast planted European beachgrass instead of our native Dune grass.

 “It is our native grass,” responded Rebecca, “but it does not build dunes.”  She pointed to a patch of American dune grass growing just along the trail. “It doesn’t grow very densely,” she explained.  Each stem of American dune grass was spaced out a bit from the others. In contrast, European beachgrass is “super dense” allowing it to better capture sand and trap it in place.

“Look, these three stems are all the same plant,” Rebecca pointed out.

“It is amazing when you start to dig down,” explained Rebecca, referring the propagation of beach grasses. Even the native “dune” grass spreads via underground stems called rhizomes; they just send shoots up at longer intervals. It is the underground growth that truly makes European beachgrass a great dune builder.

Not Exactly Invasive 

“Would it be considered invasive?” I asked at one point as we walked along, referring to the European beachgrass’ ability to spread and compete.

To be considered invasive, Rebecca explained, a species needs to meet certain criteria. “It has to do significant damage to the environment, human health, or the economy.” For example, “Scotch broom is a listed invasive weed,” said Rebecca. “It is causing a lot of economic damage.”

European beachgrass is not listed, because though it has some of the traits of an invasive species, including negative ecological impacts, it doesn’t meet the criteria.  It benefits people in many ways, so it is difficult to say it is “causing harm,” even when some native species are being impacted.

Structure of a Dune

At this point, Rebecca and I hit a junction, and we agreed to cut down to the beach. But before we headed down, Rebecca explained a bit about the structure of dunes.

“We are on the crest of the foredune; the dune closest to the ocean,” said Rebecca standing at the highest point of the dune we had been walking behind. Behind us is the backdune and looking out toward the ocean is the toe, which slopes down toward the ocean. There can be many waves of dunes behind the foredune, rising and falling just like the ocean, “and they are all covered in grass.”

Views looking out over the backdune to the ocean

Varied Vegetation

“Backdune to toe, the density of grasses will change,” said Rebecca. And you can see other differences in other vegetation as well.  Where we stood near the crest of the dune, we could see pearly everlasting, beach pea, native strawberry, and what was dubbed a “fun little thistle-y thing” growing between the blades of grass on the sand. 

Not only that but, the species of beachgrass growing in an area also influences plant biodiversity. You see, in addition, to European beachgrass, American beachgrass was also introduced to the Pacific Northwest Coast to stabilize and protect coastal communities. However, for whatever reason, European beachgrass was introduced in the south and American beachgrass in the north, creating different beach grass ecosystems. 

The differences in biodiversity between the American and European beachgrass systems is something that Sally Hacker’s Lab has studied in the past. When the lab compared the biodiversity of these dune systems, it was found that European beachgrass supports greater plant biodiversity than American beachgrass.

Two Ammophila Species

There are many additional differences between American beachgrass (Ammophila breviligulata) and European beachgrass (Ammophila arenaria) that make each type unique.

A. arenaria also has stems that are skinnier, leaves that are thinner, and a long ligule. The grass blades grow densely together; instead of sending out lots of lateral shoots, A. arenaria grows more vertically.  In addition, A. arenaria “like to be in one clump together,” said Rebecca, allowing them to “accrete more sand” and build tall, steep dunes.

On the other hand, A. breviligulata has thicker leaves and stems and a short ligule. The grass blades grow further apart; instead of growing tall, A. breviligulata sends out lateral shoots and grows horizontally. According to Rebecca, a dune field of A. breviligulata is large and “hummocky.”

Hybridization

Though historically European beachgrass was dominant primarily in the South and American beachgrass in the North, on the Central Oregon Coast, where Rebecca and I were hiking, there is a point of overlap between the grass’s ranges. And, thus, an opportunity for hybridization, or the production of a genetic cross, between the two species.

“Part of my research is about the hybrid between these two beachgrasses,” said Rebecca, as she directed me down the toe and onto the sandy beach below in search of the first little hybrid patch she has been tracking. Rebecca explained that the first time the hybrid was “discovered” was in 2012.

“Discovered is a confusing word,” proclaimed Rebecca. “Some people from my lab were doing a survey and they found grass that looked sort of weird,” but it wasn’t until later, after finding more patches of the “weird” grass before it drew much interest. And later still before Rebecca was able to do the genetic work to confirm that what was discovered was a hybrid.

A patch of European beach grass

Taken for a Ride

We combed the beach looking for the “weird” hybrid.  However, instead of finding it. we noticed many signs of storm damage—large swaths of sand torn from along the toe of the dune. 

Rebecca walked over to one of the dunes and brushed away the sand from the base of one of the clusters of beachgrass. “Look here! You can see what it looks like under the sand,” she remarked holding a dense cluster of fibrous roots that branched out in all directions, connecting to other clusters of roots by underground stems.

“They erode from the beaches,” she said, “and it gets picked up by a wave and then gets spread from one beach to another.”

Rebecca removing sand to expose the underground root and stem system beach grass uses to propagate

The Path Back

We continued looking along the beach, passing by the spot Rebecca recalled finding the hybrid in the past. The patch had since been washed away, perhaps to start up someplace new.

We headed inland toward another area Rebecca had GPS coordinates for, chatting along the way about her love of plants and how curiosity regarding invasive species led her to her work today. “Why do some plants get introduced and nothing happens?” Rebecca wondered aloud, reminiscing.

Rebecca’s passion for research was evident as we talked further. She told me how she got into genetic work in order to obtain higher resolution data that would allow her work to have a larger impact than in the past. “Genetic work is 100s to 1000s of data points,” she marveled.

Ultimately, Rebecca’s careful consideration of her interests and skills landed her in Sally Hacker’s Lab.

The Hybrid Problem

And then we were there, standing in front of a large path of hybrid beachgrass. “Here is its sweet little intermediate ligule,” Rebecca smiled as she pulled the leaf down on one of the grasses to reveal its “pit” tissue. “This is one of our wonderful patches,” she remarked, her gaze sweeping down at the bunches of grass that grew at our feet.

It was at this point that I asked Rebecca about the implication of the hybrid. “Is the hybrid a problem?”  I questioned.

Even though they are not native to Oregon, explained Rebecca, beachgrasses are providing a service to people that live in coastal areas by protecting them from storm surges and erosion.  However, “the two species produce different shaped dunes that have different value,” Rebecca explained. If the dominant species changes, that could mean a higher risk of storm surge to overtop the dunes, or it might not. She cited a 2010 or 2012 study that conversion to American beachgrass in areas where European dominates would create “a three-fold increase in overtopping risk.”

When it comes to the hybrid, there are a lot of questions to answer.

“What are its impacts going to be on the dunes? Is it going to build different shaped dunes? Is it going to take over areas of parent species?” Rebecca listed. “It’s kind of a dash to figure out the impact.”

A close-up of the short ligule of the American beachgrass.

More Invasive

There are also questions about how the hybridization might increase the invasiveness of the beachgrass species, potentially harming native species to a greater extent than the parent species.

“Hybridization can jump-start evolution in plants,” explained Rebecca. When separate species are put together you get something brand new— “a completely novel genotype!”

In addition, there is some evidence that the hybrid may be able to produce viable offspring, allowing for even more crossing of populations, increasing genetic biodiversity further. “Increasing genetic variation in invasive plants has been shown to increases their invasiveness,” said Rebecca.

To be Done

So, what can be done? “Can we stop or prevent the spread of invasive species?” I asked Rebecca, as we stood considering the potential impacts of her newly identified cross.

“There are a lot of layers to think about,” Rebecca suggested. Questions regarding who is being impacted and how will help determine an appropriate course of action. She used the example of cheatgrass, pervasive invasive grass in rangeland environments. The impacts of cheatgrass affect ranches and those that depend on the work the ranchers are doing, as well as the native plants and animals that live in the ecosystem.  Thus, figuring out how to prevent the spread of cheatgrass is a high priority in the west.

With beachgrasses, it is a bit different. Again, despite some of the ecological ramifications, the presence of the grasses is needed in human-inhabited coastal areas as protection.

However, there are some efforts to remove the beachgrasses and restore some of the native habitats that have been virtually eliminated on the Pacific Northwest Coast with the introduction of the grasses. In fact, Rebecca shared how her advisor, Sally Hacker, is involved in some research looking into how to best restore some areas of the dunes that aren’t critical to coastal protection.

“When you think about the impact,” said Rebecca, “we should all be on board to keep invasive species in their lane.” 

Three of a Kind

At this point, Rebecca and I headed back down the path toward our vehicles. Then, looking out in the distance, Rebecca pointed out a large dune that sat just in front of a line of houses.

“It is covered in hybrid,” she said, “and has all three species.”

Rebecca went on to explain how, instead of a natural system, the dunes in the area are actively managed—built up and planted by people—and suggested that we take a quick look. So, when we got back to the trailhead, a short walk later, we both hopped in our vehicles, and I followed her over to see what the fuss was about.

Upon arrival, we were able to find our first sample of American beachgrass with its short ligule. We also saw more European beachgrass, some American dune grass, and, of course, a lot of the hybrid.

Rebecca told me that so far, they have found more than 27 patches of the hybrid at 17 different sites heading North from Pacific City. The site we were visiting was the southern terminus of its extent so far.

And there was a lot of it! Rebecca pointed out the patches as we walked back onto the dune and looked around. There is so much of it, that Rebecca told me that they were in the process of training people to identify the hybrid as part of a Citizen Science Project.

We didn’t stay long on the managed dune, but Rebecca helped me gather a sample of the European beachgrass, American beachgrass, and the Hybrid. We lined them up and snapped a picture of all three for comparison.

European beachgrass, American beachgrass, and the Hybrid

A Bleak or Bright Future?

Looking at the photo now, with the hybrid placed alongside its parent species, I feel a bit like I am looking at a photo of a high school graduate with their proud parents.

We know a good deal about the European beachgrass and American beachgrass, they are settled in the habits of building dunes in the same way they always have, but the hybrid is something new—a fresh mix of genes with so much potential. The future of the hybrid could very well change the Pacific Northwest Coastline, for better or worse.  Many questions remain. And a lot of opportunities await. Fortunately, we have people like Rebecca here to help us understand its future.

Rebecca Mostow is a Graduate Fellow at Oregon State University in Sally Hackers Lab studying dune ecology. She has also worked as an environmental educator and research technician. She earned a Bachelor of Arts in Biology from Oberlin College in 2013.

Hike with a Wildfire Ecologist

Views from the open ridge top on Sterling Mine Ditch trail.

Fire. Red, hot whirls of gases set ablaze in the presence of oxygen. It is beautiful. It is dramatic. And it is dangerous. Right?

When I think about fire in my own life, phrases like “don’t play with matches” and “only you can prevent forest fires” spring to mind.  I imagine firefighters battling blazes, blackened trees, and billows of smoke filling the sky— in short, destruction.

This view of the fire is not uncommon. Over the years, fire has developed a bad rap. Fires have wreaked havoc on millions of acres of forests in the west. Thousands of people are displaced and even die each year as a result of fire. It has been suppressed and fought against for a good part of the last century.

We fight fire. But should we? 

After spending the afternoon talking and exploring the hillsides of Southwest Oregon with Chris Adlam, OSU extension forester, I am inclined to say “no.” Or at least, “not always” and “it’s complicated.”

Chris Adlam stopping for a photo on the trail.

The Hike

  • Trailhead: Bear Gulch Trailhead, Sterling Mine Ditch Trail
  • Distance: Approximately 4.7 miles loop with longer options.
  • Elevation Gain: Approximately 690 ft
  • Details: No restroom at the trailhead and limited parking.

Welcome to California

I met up with Chris on a warm spring afternoon at the Bear Gulch trailhead. Moisture hung in the air teasing us with the prospect of rain, but except for a sprinkle or two, we stayed dry. Dry is a common condition in the part of Oregon, which, as Chris expressed, is “much more common in California.” He described the place we were hiking as “oak woodland,” an ecosystem characterized by its hot, dry summers and little rain.

“As far as the ecology is concerned,” said Chris, “we are in California.”

Oak woodland ecosystem at the start of the hike.

It’s Complicated

Of course, there is one more thing that defines an oak ecosystem—fire! “This is a landscape that was made this way by fire,” explained Chris. But, as we started up the trail, it became evident that this defining characteristic had not been present on the landscape for a long time. 

“It looks messy,” said Chris pointing to patches of dead manzanita with new growth on top. There were also a lot of trees, including conifers crowded together along the path. “It looks terrible to my eyes,” Chris exclaimed.

Chris told me how he had found an old snag not to far from where we were hiking that showed “17 fire scars.” He explained that these scars mark the frequency of fire in the area. “It would have burned every 3-5 years,” said Chris. But that was the past. Looking around at the crowded hillside, Chris frowned. “This place hasn’t seen fire in decades.”

An Education

We moved up the trail slowly, taking in the scenery and flowers. Both of us botany nerds, we paused frequently during the hike to marvel at the rich plant diversity we encountered. After attempting to take photos of Henderson’s Shooting Star, I asked Chris to tell me a bit about his background.

“Well, my family growing up spent a lot of time outdoors,” Chris began. He recalled going birding and visiting the mountains. But “I never thought that much about it,” said Chris. Later, he became interested in sustainability and ecological restoration. “You can help fix them,” Chris said, referring to ecologically damaged areas, “And I thought that was very powerful.”

Mindshift

Chris grew up in France and moved to Canada when he was 19 years old. His earliest understandings of Native people had come from John Wayne movies. It was not until he moved to Canada, where “the tribes were very visible,” that he has his first encounters with Native people. He saw Native people fishing in rivers and interacting with the landscape. He was intrigued by their connection to the land.

Chris, like many from a western culture, had the perception that people are separate from nature. That you “go to nature,” rather than being a part of it. Stemming from this is the idea, he was of the mindset that “people necessarily hurt nature”—a contentious relationship.

However, Chris learned during his time in Canada and later in the United States, this does not have to be the case.

“There are many cultures that live sustainability not by separating themselves from nature but taking responsibility for their role as caretakers of their ecosystem.” It was this shift in thinking that really got Chris interested in fire ecology.

Cultural Problem

As we gradually reached the ridgetop, Chris spoke more about fire and how his interactions with Native people taught him to appreciate fire.

“I have been fortunate to learn from many tribes—the Karuk, Yurok Tribe, and Confederated Tribes of Grand Ronde…” began Chris.

“What have you learned?” I asked.

“I think about it a lot,” was his response. “Most important to me,” he concluded after much consideration, “is understanding how Native people treat fire differently.”

“We tried to fight fire; we tried to prevent it,” said Chris, but “you can’t control fire!”Chris went on to explain that the “fire crisis,” as he put it, “is a cultural issue.” “It is not a lack of science or knowledge. Not a policy support problem. It is a culture problem.”

“For Native people, fire is a sentient living thing.” While in the United States, it is legally considered a nuisance if it is not controlled.

FireKeepers

For centuries Indigenous people worked alongside fire to actively manage the land. Fires were set regularly before European settlement to maintain an open landscape ideal for hunting and foraging.

Walking along the ridgetop now, though, overgrown with vegetation and littered with deadwood, it was hard to imagine the open prairie ecosystems of yore. The views would have been incredible!

Botanizing

Despite the overgrowth, there were still some lovely remnants of the past landscape to admire, though I think Chris had a hard time getting past the devastation.

In particular, Chris and I admired the native wildflowers along the trail. Chris had a knack for identifying plant species

Many of the plants we observed were used by Indigenous people as food, fuel, or fiber. I asked Chris to point some out. He pointed to a fuzzy white flower with rounded petals that pointed at the tip—the Cat’s Ears Lily. “These have an edible root,” said Chris. Traditionally, they would have been used by a lot of tribes as a food source, he explained.

We also saw several examples of yellow Lomatia or biscuitroot. I had heard of the use of Lomatium to produce a flour that was shaped into transportable cakes. In addition to having edible roots, Lomatium has edible leaves and flowers, Chris told me.

“And then, of course, are the oaks!” exclaimed Chris. The acorns of the oaks were staple foods for many tribes who, at the time, lived in an oak-dominated environment. Once the dominant tree across the landscape, many tribes depended on and managed for, oaks and their associated species.

And what was used to manage the land for these plants? You guessed it. Fire.

A Cat’s Ear Lily. One of many wildflowers found on the trail.

Fire Dependent

Manzanita is another species used by Indigenous people found on our hike. “They made the berries into a drink,” said Chris.

At this point, we walked past a tall manzanita tree. Though beautiful to the eye, “ecologically they are an aberration,” said Chris.

Like other important indigenous plants, manzanita is a fire-dependent species. They need hot fire for seeds to germinate and plants to grow. Looking around the area, there was no young manzanita to be found. Instead, the shrubs were old and dying.

Manzanita was found all along the trail.

Ceremonies with Fire

Fire was also used in many tribal ceremonies.  Unfortunately, a lot of ceremonies were made illegal, Chris told me, “even until the nineties!” So, ceremonies were practiced in small groups or families. I asked Chris if he was able to observe any of these practices. He said that there are some you can go to but didn’t speak to his own participation.

Instead, Chris told me about one of the ceremonies observed by Indigenous people in Northern California until the early 1900s. The ceremony took place at the landscape level. One mountainside was set ablaze, and then another, and another. Eventually, the signal reaches the medicine man who is waiting in the valley bottom to “call the salmon” home.

At the same time, the landscape responds. As the shrubs and grasses burn on the mountainside, the water level is raised in the river. The smoke from the burn fills the valley, blocking the sun’s rays and cooling the water. All of these help the salmon return. “It is more than a lot of symbolism,” Chris proclaimed. 

Prescribed Burn

With all this in mind, Chris made the argument that we need to burn. And a lot more than we are doing right now. By doing this, we will not only revive the landscape but learn to appreciate fire and dismantle our need to control it.

“The most important thing to me is to do more prescribed burns. If we can figure out how to do that, then we will be good,” Chris stated.

A prescribed burn is a controlled process where landscapes are set on fire systematically. Prescribed burns are started from the top down. The landscape is lit with torches in lines and allowed to burn downslope until the fire burns out, then another line is placed below it and allowed to burn.

The whole process is easy according to Chris. “Regular people do prescribed burns,” explained Chris. “That is how we have done it over this last century.”

And prescribed burning has been used virtually everywhere across the globe. “You can go anywhere, people use fire.”

It wasn’t until recently that prescribed burning has, as Chris put it, “been given to the professional elite.” This change in responsibility is a mistake, according to Chris who has made it a goal to get people who live in fire-adapted landscapes involved in the process. “Everyone who lives in a place like this should get a chance to go to a prescribed burn,” he stated. It is part of our history and needs to be part of our future.

In Favor

We passed by another area filled with dead and down wood.  I started wondering why we were not burning more landscapes and I asked Chris about it.

“For a long time, we thought we could stop fire,” suggested Chris, but recently, “I think it is changing.” People are beginning to understand that prescribed burns are okay. In fact, according to Chris, eight out of ten people nationwide are in favor of the practice. People are getting it. They are understanding we need it. “It is not controversial,” said Chris, as some might make you think.

Overall, “we need to get people together to talk about it,” Chris suggested. “We need to agree. We can’t fight about this.” Being flexible and looking at the management options available, including prescribed burns, is key.

Several areas next to the trail were covered with dead and down wood.

Doing Nothing is Deadly

Chris and I reached a viewpoint or at least a partial one. For the first time in a while, we could see out onto the adjacent hillsides. “This place should be oaks and a few madrones,” said Chris. “We should see all the way out.” Instead, spindly orange conifers, sick from bark beetle infestation, dotted the crowded hillside.

A partial viewpoint. Looking out we would see many dying conifer trees in the distance.

“ These trees are a bunch of garbage and are doing to die !” Chris exclaimed looking around us. He pointed to a sickly sapling that was losing foliage and a tall Douglas-fir riddled with cones. “That one is dying,” he stated matter-of-factly.  Both losses in upper crown foliage and producing a lot of cones are signs of stress, Chris explained.

“If a wildfire came through it would do great things ecologically,” Chris continued. The area would be cleared of much of the sickly Douglas-firs, which would allow the oak tree to flourish. The fire would help any remaining trees better able to fend off insect attracts.  The manzanita and ceanothus would be able to go to seed. Grasslands would be reestablished.

 As I walked the ridgeline, I tried to imagine it—this place 100 years ago. I tried to see through Chris’ eyes what was lost and what one day may be found.

A stressed Douglas-fir tree with many cones.

Fire Scars

Eventually, we passed an old snag, cracked and open so you could see inside.  Chris later called it a “cat-face.”

“This one had fire,” said Chris as he crouched down to get a closer look. Tracing his fingers across the black markings that punctuated its open face. “There are at least six here,” said Chris, referring to the multiple fire scars found on this one tree.

Each time a tree is burned it leaves a blackened mark. Given time, the tree will try to grow over the blackened area. However, like in the snag we found, it does not necessarily heal completely over, so that sometimes fire scars remain exposed for years to come.

The fire scars on this tree were close together, Chris estimated that the tree had been scared 3-4 times over a 20-year period. “Not to mention those that didn’t make a mark!” Chris clarified.

A fire-scarred snag, evidence of fires in the past.

Legacy of Fire

As we continue along, Chris pointed out more examples of the legacy of fire on the landscape:

We saw another “cat-faced” tree—an oak—but this time its face was completely healed over so you could not see inside.

There was a notable area that was mostly clear of trees and underbrush, except for a few oaks with spreading branches. Though it was difficult to say for certain what had cleared the area, Chris had a sneaking suspicion it was fire.  “I would guess 1987,” said Chris, as a fire had been in the area that year.

“The oaks are doing great,” Chris pointed out as a testimony to the efficacy of fire. An understory of roamers fescue, a native bunch grass, punctuated scene.

An area cleared of underbrush and native grasses growing.

We also saw countless madrone on the trail, some that were huge having survived the fires of the past.  Others with multiple stems, creating a tentacle-like effect that was mesmerizing. Chris explained that almost all broadleaf trees, including madrone, will send out multiple shoots following a burn. Though some trees, like oaks, will thin themselves out, madrones keep their multiple stems. These octopus-like trees were a result of fire!

Madrone tree with many trunks.

Legacy in the Landscape

Chris also noted how different areas within an ecosystem will have different legacies of fire. “South facing slopes tend to burn hotter and will have more oaks and fewer conifers, while cooler north-facing slopes will not burn as hot and have more conifers, Chris explained. In general, all areas “burn more severely on the upper part of slopes,” said Chris.

As we walked down into a stream drainage, Chris continued to ruminate on the topic. He speculated that the drainage would offer a different outcome when it comes to fire, perhaps acting as a fire stop and protecting the area from a severe burn. However, he prefaced that the steepness of the slope could also help spread the fire more quickly if it got started lower in the drainage.

Essentially, by looking at the contours of the land and the ecosystems that exist there, you can start to piece together a fire story.

Oh, the Plant Biodiversity!

Okay, so we talked a lot about fire on your hike! But another topic that came up time and time again was plants! “I aspire to know every plant where I live,” he admitted at one point. 

Considering I also have a strong affinity for plants, Chris and I spent a good amount of time distracted by the botany all around us.

We saw everything from milk vetch with its fuzzy seedpods to balsamroot with its bright yellow flowers. We saw paintbrush, wild carrot, sweet cicely, fern leaf biscuit root, nine leaf biscuit root, Henderson’s fawn lily, and larkspur, to name a few. Fragile fern was another hit along the trail. The biodiversity was amazing!

And the enthusiasm of Chris equally so. At one point, Chris charged off the trail to check out what he thought might be a cypress tree, only to find out it was a juniper. We were seeing species that you find in western Oregon, co-mingling with species from eastern Oregon.

It seemed like around every bend, was always another fantastic botanical find!

Of course, Chris could not help himself. “People don’t always think about it,” he said as we passed by another patch of manzanita, “It is fire that creates all this diversity.”

A Ceanothus bush in bloom.

Not Clear Cut

As we came back up from the drainage, our views out became less obstructed than earlier in the hike. We could see hillsides of grasslands with ceanothus patches rising all around us. However, that is not all we could see. Off in the distance was a large patch of open ground—a clear cut.

Chris stopped in his tracks. You could tell he was not happy about what he was seeing. Though he admitted that timber harvesting was not his area of expertise, he saw some huge issues with how it was being done in Oregon. “Oregon has the least restrictive laws,” he started in on the issue. “You can clear cut 120 acres and if you throw a stick back in, you are all good.”

You could sense Chris’ frustration, which he tempered quickly. “There is a place for it,” he admitted, referring to Oregon’s coast range. It can provide “important habitat for bird species,” he went on, but this was not the place for it. “Dry forest is not good for this.”

Then there is the concern for fire. “Plantations burn more severely than older forest,” Chris said. Once the trees start to regenerate, all the benefit of fire protection that might come immediately following a clear cut is gone. “5 years you don’t have to worry, but once it is tall again, it will carry fire.”

Out in the Open

As we moved further along the trail and our views continued to improve until ultimately, we found ourselves in an open prairie ecosystem—face to face with the past. “I think this is a lot more what the landscape would have looked like,” said Chris.

Stately Black Oaks were spaced out at irregular intervals on the windswept hillside. It was beautiful. And it was the result of centuries-long practices of using fire.

Chris talked about how Indigenous tribes not only burned to maintain the open meadows ideal for hunting deer but how it was also used for pest management. For example, burnings occurred at night so moths that infest acorns would be attracted to the light and burn up.

A black oak tree in the open prarie.

Invasive Grasses

Amongst the fields of native grasses, there were also many invasive grasses, like medusahead, vying for space on our green hillside. Chris explained that a lot of the invasive grasses are annuals that go to seed each year, repopulate, and spread. The native grasses, on the other hand, are perennials that don’t seed every year, and thus are better adapted to frequent fire.   

Without frequent fire, the invasive grasses were encroaching on the lands of the native grasses—an all too familiar tale. But, as ever, Chris had the remedy—fire.  He explained that the annuals could be killed off with a well-timed prescribed burn, leaving the perennial natives to thrive.

The trail winding its way through the open prairie.

Working toward Change

Despite the presence of invasives, Chris was in his “happy place” as we walked along the rolling hillside.

Chris and I talked about his work as we went. “So far I am working with landowners,” he said, helping them go through the steps of doing a prescribed burn. He also hopes to work with tribes in the area to bring fire back to the landscape.

Eventually, we reached the opening to a tunnel and a sign that read “Turning Water into Gold.” It was the opening to the Sterling Mine Ditch Tunnel—a hand-dug, 26.5-mile tunnel used to divert water for gold mining.  After poking around for a few minutes, we continued our conversation.

Sterling mine Ditch Tunnel

Changing Minds

“I want to bring people to more burn areas too,” remarked Chris, we began descending the trail through some oak woodlands and back to the road. He explained how people that experience fire are often traumatized. They cannot imagine the burned area ever being beautiful again. Chris wants to break that cycle of trauma by bringing people to a burn site so they can see “that not all is lost.”

Following a fire, life remains. There are still many live trees. Water flows through the streams. An entire seed bank of herbaceous plants, including brightly colored wildflowers, awaken and bloom.  Animals, like woodpeckers, reptiles, deer, elk, and bears thrive in a burnt landscape.

Chris talked about using nature journaling as a tool for self-reflection on the burnt landscape. “It is a project I am working on,” he said.  Nature journaling involved getting people to record what they notice in pictures, words, and numbers, generating questions, and reflecting on their own experiences.

By experiencing firsthand, the resiliency of the landscape to fire, Chris hopes to get people in touch with their own resilience. Experiencing “hard things” does not necessarily equate to a diminished existence, Chris explained. “You can come up all the better.”

Slow Burn

Chris and I continued to chat as we descended downhill through the woodlands to the gravel road and then back to our cars.  Upon reaching the road, we said our goodbyes.

I really enjoyed my time talking and walking with Chris. His passion for his work with fire and its ecological and cultural significance was evident throughout our time together. 

Since then, the idea that fire is a force to be valued and respected has been slowly burning in my consciousness.  To think of fire as something vital is a paradigm shift, but one that makes a lot of sense both ecologically and socially. It is easy to become alarmed by the fires that burn through the west each summer, but if we can start to see fire differently, we can learn to respond to fire more appropriately and learn to adapt.

Chris Adlam is the Regional Wildland Fire Specialist for Jacksonville County Oregon State University Extension. Chris completed his Ph.D. in Ecology with a focus on revitalizing the use of fire in managing the land with northern California tribes.


Hike with Urban Wildlife Ecologists

Sam and Yasmine own by the water at low tide

It was an unusually sunny spring day in Seattle when I arrived at Discovery Park to hike with Sam and Yasmine—two energetic, young urban wildlife ecologists from the University of Washington. I was a bit early for our meeting, so I decided to wander down one of the many trails and do a little exploring.

Native trees and shrubs lined the trail, wildflowers were in bloom, and bird song filled the air. I watched a white-crowned sparrow hop from shrub to shrub and branch to branch, as light filtered through the canopy. It was a peaceful and pleasant ramble. You could almost get lost in nature’s spectacle if it were not for the other visitors that shuffled by at regular intervals. 

That is the thing with urban parks, they are sort of a mixed bag—both a respite for wildlife and a central hub of activity for the populous. Often, they are the only way many people can access wild space. But just how wild are these spaces? And what becomes of the wildlife that call the “urban jungle” home?

Shortly after returning to our meeting spot in front of the visitor center, Sam joined me with her dog Sequoia, in tow. Yasmine arrived only a few minutes later. Quick introductions and an exchange of M&Ms between friends, and we were off on the trail.

The Hike

  • Trailhead: Discovery Park Visitor Center
  • Distance: approximately 3 miles (12 miles of trails)
  • Elevation Gain: unknown (varies)
  • Details: There is ample parking at the trailhead and several routes to choose from. When the visitor center is open it has restrooms, informational displays, and maps.

Youthful Indiscretions

Both Sam and Yasmine grew up in urban areas—Sam outside of San Francisco and Yasmine just outside of DC.  During their youth they also both spent a lot of time outdoors.

“I was always obsessed with animals and being outside,” said Sam about her childhood. “I was that nerdy kid in the classroom reading animal encyclopedias….” she went on.

“I got lyme disease a couple of times because I was always running out into the woods,” Yasmine shared.

Urban Wildlife Ecologist

Both were also pulled toward urban wildlife.

Sam recalled the area she grew up— “I was amazed by how much wildlife is there.” This realization coupled with opportunities to get involved in research at the undergraduate level helped direct her academic future.

Now Sam is in her first year as a Ph.D. student at the University of Washington studying Seattle’s coyote population.

“I look for coyote scat,” she said bluntly. “There are supposedly coyotes all over Seattle, but I have yet to see one. I have found their poop in some places through.”

Yasmine, on the other hand, started her academic career pursuing vet school. Like Sam, however, she got involved in undergraduate research—studying invasive fish in the Chesapeake watershed.  She worked on a variety of projects but ultimately kept coming back to urban wildlife.

Yasmine is now in the early stages of her Ph.D. program. “I am still figuring it out,” she explained. “I am going to be collecting carcasses to look at urban wildlife health,” she went on—to look for parasites, viruses, and assess their overall condition. She plans to source coyote carcasses from the USDA and Washington Department of Fish and Wildlife’s Control Operations.  As she put it—she will be “recycling” carcasses—giving them a new purpose.

Human Discovery

I followed Sam and Yasmine along a well-established trail, passing by both native plantings, grassy knolls, and large swaths of invasive species. While we walked, Yasmine shared a bit about Discovery Park’s history.

“It was used by a lot of tribes for thousands of years,” she began.

“And then it became a military base, and all of this became raised for horse pastures and hundreds of buildings at its prime, like 80 years ago,” Yasmine gestured around. Finally, in the 1970s it was repurposed as a park and restoration work became.

It is an “earlier succession park,” according to Yasmine, as it was planted only in the last 50 years and it is still undergoing active restoration. 

Sam piped in with the size of the park—”534 acres.” That is a nice chunk of real estate for urban wildlife. 

Lots of invasive species and native planting at Discovery Park

Scoop the Poop

As we continued past a few buildings and through pockets of forest along the trail, I asked Sam to elaborate on the scat project.

“I have only collected a couple of scat,” said Sam. “Our main push for scat collection will be this summer.” She explained how there was some concern about the quality of scat collected in the winter with all the rainfall in Seattle. But she was able to get some good quality data from the few she collected.

Now you might be thinking, why scat? Why study excrement? Choosing to study something like scat, begs the question— “Why?” I asked Sam 

Sam explained how her work is in collaboration with Woodland Park Zoo and Robert Long who is a proponent for non-invasive carnivore survey techniques. You don’t have to handle an animal to learn all about it. The idea of non-invasive techniques is that you can learn a lot without interacting directly with the animals you are studying, thus reducing potential unintended stress or harm. “I think it is really cool,” Sam exclaimed.

Plus, there is a lot to learn! Sam swabs the outside of the scat to “identify the coyote that pooped the poop,” as she put it most eloquently. Then the inside of the scat sample is swabbed and analyzed to determine what it consumed. “I will be doing everything genetically,” said Sam. However, according to Sam, you can determine a lot about the animals’ diet by manually going through it and looking at the hairs embedded in the undigested remains.

Why care?

Our eyes peeled to the ground, Sam, Yasmine, and I continued to scout the area, dreams of big piles of poo dancing through our heads. As we walked, I asked Sam and Yasmine to tell me more about urban wildlife and why someone should care about keeping tabs on the urban jungle.

Yasmine spoke up first “It is beneficial for us and to them to learn how to coexist,” she stated because “they are here anyway.”

If we understand urban wildlife better, we can learn how to respond to their presence and develop management techniques that make sense.

She went onto discuss, as she put it, “the disease angle.” “A lot of these animals are vectors for disease,” she explained. “How can we ensure they have enough space, so they don’t end up in human spaces?” Therefore, it is important to understand what makes urban wildlife tick; “so, they don’t pass a disease on to our pets or kids or something.”

Clever Coyotes

Even if people wanted to eliminate coyotes from urban environments, which I believe Sam and Yasmine would argue is a mistake, it would be very difficult to accomplish.

“They are very adaptable,” shared Sam. “When you remove coyotes it creates a vacuum that coyotes will go fill… They have density-dependent fecundity.” Meaning, if you reduce the populations, coyotes simply produce more offspring.

Yasmine agreed in Sam’s assessment. “They really thrive in so many different cities in ways other animals don’t,” said Yasmine. Sharing how during the first coyote project she worked on she found coyote using railroads and living in trainyards. It “blew my mind,” said Yasmine.

Cool Coyotes

After walking past several viewpoints along the trail, I asked Sam and Yasmine if they knew of any other “cool coyote facts?”

“They are the top predator in Seattle,” responded Yasmine. They suppress many other meso-carnivore species, like skunks, raccoons, opossum, and foxes.  This can, in turn, boost overall biodiversity and ecosystem functioning by allowing prey of smaller predatory species to survive.

And what about people? The jury is still out. Yasmine explained that you will find papers saying opposite things when it comes to how coyotes respond to people. “It seems to vary by city,” said Yasmine. In some places, it seems they avoid people spatially, while others say they don’t mind being in the same space but will avoid people temporally. Either way, they don’t like us very much.

Of course, there is still much to learn about coyote and how they interact within their community. They are part of a “messy web,” said Yasmine.

A nice viewpoint next to an area being restored

Coyote Threats

So, with everything seemingly going well for coyote in urban environments, I asked Sam and Yasmine if there are any threats that coyote face.  Yasmine had mentioned parasites and diseases as part of her research project. Is there something out there wiping out coyote populations?

The short answer seems to be no. But Yasmine did share a few threats that coyotes face.

“The first thing that comes to mind is mange,” responded Yasmine. Caused by parasitic mites, mange is a problem for coyotes that live in colder climates. Infected animals will scratch themselves too much, so that they lose their fur, leaving them susceptible to the elements.   

Environmental toxins are another challenge. Led and arsenic are also potentially problematic to coyotes. As well as anticoagulant rodenticides. These chemicals have the potential to bioaccumulate or build-up, in the tissue of animals. They can also be biomagnified (increase) through the food chain, such that predators, like coyotes, face the brunt of the toxic effects as they consume prey riddled with toxins.

Food for Thought

At this point, we followed the road down to the beach. We passed by stands of stinging nettle and Yasmine shared her favorite ways to harvest and prepare stinging nettle by blanching and sautéing it. All this talk of food, of course, got me interested in learning more about Sam’s project.

“What do coyote eat?” I asked.

Though DNA analysis of the collected scat has not started, Sam told me that there is a lot that can be discovered by simply looking at the scat. As far as Sam has seen from samples found at one site, coyotes are eating rabbits and snakes, but also candy bars. Just like bone and fur are preserved in the scat of coyotes, so are wrappers and other pieces of plastic.

Additionally, a good deal of research has already been done on the coyote diet. And findings are incredibly variable. Sam explained that what coyote eat “depends on where you are and the time of year.” At one location in the North East, for example, the coyote diet was “80% berries at one point,” said Sam. Cities teeming with black rats, roof rats, and eastern cottontail are prevalent sources of food—all invasive species.

Finally, Sam said that household cats do not appear to be a regular part of coyote diets, despite what some would believe. Though one site in Los Angeles may be an exception.

Scoop the Poop Reprise

I was not “dung” with this line of questioning, however, and I asked Sam if she knew from her research how many coyotes inhabit the Seattle area?

Though she didn’t know offhand, Sam shared how the scat she was gathering—using a technique called “mark and recapture”—could also be used to determine population size.

The outer coating on each sample of scat contains epithelial gut cells that can be genetically identified down to an individual. As Sam put it, we know “exactly which individual pooped the poop.” With enough sampling, some individuals are likely to be “recaptured,” or identified a second time. It is the recapture data along with the initial captures that allow scientists to estimate population size.

This begs the question—how hard is it to collect samples? According to Sam, it is as easy as picking up your own pet’s waste—only she uses two Ziplock bags while collecting.

Down by the Sea

As we marveled at the amazing advancements in DNA analysis, Sam, Yasmine, and I made our way down to the beach. Sam’s dog Sequoia led the way down to the water. The tide was unusually low, so we decided to walk the shoreline for a while, dodging sea anemone and other critters that lie underfoot.

We talked about grad school, tutoring, and hiking in the Pacific Northwest, among other topics, including scat, as we walked the beach.

“One of the coolest coyote scats I have ever seen was on the Strait of Juan de Fuca on the Olympic peninsula,” shared Yasmine at one point. “There was one on the rocks in the tidepool. There were crabs and mussel shells in the scat!”

Unfortunately, our beach adventure did not turn up any such gems.

Low tide at Discovery Park

Social Structure

It did, however, turn up some juicy gossip on coyote social structure. Coyotes, according to Sam, live in family groups, but often act independently. “It is thought to be one of the reasons for their success,” explained Sam. The flexible groupings allow them to hunt in groups when it is advantageous, or head out on solo or couple adventures.

In addition, coyotes maintain territories that vary in size depending on how much food is available. Territories are defended by members of the family groups or packs.

Coyote Careers

After soaking in some sun, Sam, Yasmine, and I headed back uphill to continue our search for scat. We entered a large field/lawn area that looked promising. But sadly, our efforts were not rewarded. Still, no sign of coyote, though we did hear sea lions barking in the distance.

Feeling a bit defeated, but still hopeful, we continued uphill, our senses on high alert. I wondered what Sam and Yasmine felt about their chosen line of research. I asked what advice they might give to the next generation of wildlife ecologists.

Sam was first to respond. She explained how working in wildlife ecology is “not like what you see in National Geographic.” A lot of opportunities to study wildlife ecology are non-invasive and local. Many international jobs can be exploitive and/or very competitive. She recommended: “Know what you are interested in ecologically.”

Yasmine added, “You need to be flexible…build your own way.”

“More and more the field is becoming collaborative,” added Sam. “Getting involved in projects is a good start.”

Yasmin and Sam pose for the camera with Sam’s dog Sequoia

Citizen Science

Getting involved does not have to start with graduate school!  Citizen science projects in urban wildlife and other sciences are becoming more popular.  Since our hike, Sam and her collaborators have launched the Seattle Coyote Study website where people can sign up to help collect coyote scat for the project.  Volunteering is easy, fun, and flexible, as participants choose when they go out and how often. Check it out at seattlecoyotestudy.wix.com/seattlecoyotestudy.  

Eyes on the Prize

Around this point, we saw a used dog poop bag plopped on the side of the path. Not the sign of life we were looking for, even if it was technically scat.

Our hunt turning up nothing but domestic dog poo, I asked the duo what other signs of wildlife might be fun to look for in an urban setting. There has got to be something better than this!

“Deer sign is one the easiest things to look for,” said Yasmine. In some places, you can see a clear browse line. In other places, it is harder to detect but are still able to find signs of browse on individual plants.

Sam and Yasmine both agreed that tracks are also a lot of fun to look for, especially in the mud or snow. “When mud has that glaze,” said Sam, “it preserved prints perfectly!”

“And then if you are into birds, listening and looking for birds,” said Yasmine is a great way to connect with urban wildlife. She admitted she has never been that “into birds,” but has grown a greater appreciation for the birds in her neighborhood recently.

“It is amazing the diversity of birds in an urban area,” added Sam. “I was walking around my neighborhood and there was a pileated woodpecker on a telephone pole!” A rare site indeed!

Leaving a Mark

Eventually, Sam, Yasmine, and I made it back to the parking lot. Having completed our loop, we had not turned up a single sample of scat for Sam’s research. Defeated but not down, we said our energetic goodbyes and parted ways.

Upon reflection, though I did not find what I was looking for per se, I found something far greater. Spending time with Sam and Yasmine—their young enthusiasm for research and science—was hopeful and invigorating. There is a lot of good, thoughtful science happening, right now! It might go undetected much of the time. It might even be ignored. But like coyotes in Seattle, the signs are there. You just got to keep looking.

Samantha (Sam) Kreling and Yasmine Hentanti are both Ph.D. students studying urban wildlife at the University of Washington. Sam has a B.S. in Molecular Environmental Biology from the University of California Berkeley and Yasmine has a B.S. in Wildlife Ecology & Management from the University of Maryland.


Hike with a Predator-Prey Wildlife Ecologist

View of the forest from the trailhead.

Wildlife populations are dynamic—ever-changing, depending on circumstances. They increase when resources are abundant and decrease when resources dwindle. Populations are limited by numerous factors that affect their ability to survive and reproduce. Factors, like predation or disease, that limit high-density populations more readily, as well as more sweeping large-scale factors, like habitat change or loss.

However, limiting factors are also wide-ranging and often entangled with each other—a giant web of factors where if you pull on one thread, others are sure to respond. This is where Taylor Ganz, Ph.D. candidate from the University for Washington, comes in. Taylor has been working for the last four years to understand how deer and elk populations in Washington have been changing as they respond to factors in their environment.

I met up with Taylor on a mostly sunny day for a hike in the Capitol State Forest near Olympia, WA to talk about her research and to learn more about population dynamics.

The Hike

  • Trailhead: Mima Falls Trailhead, Olympia, Wa
  • Distance: 6.5 miles
  • Elevation Gain: 700+ feet
  • Details: There is ample parking at the trailhead and a pit toilet. Trailhead is very accessible, but Discover Pass is required for parking. There are many options for trail routes through the area. The Mima falls Loop is part of Olympia’s Capitol State Forest.

For the Love of Science and the Outdoors.

As we started down the tree lined trail, Taylor told me a bit more about her background and research.

“I have always loved science and being outside,” said Taylor. So naturally, in college, she studied physics and mechanical engineering. “I thought I would work in alternative energy design and development,” she explained. 

The problem was instead of doing engineering internships, like her classmates, she found herself drawn toward jobs in outdoor recreation and education. “I was working as a wilderness ranger… a flyfishing guide… a rock-climbing instructor…” For 5 years she worked for NOLS (National Outdoor Leadership School) before deciding “to get back to science.”

She started out at the Yale School of Forestry where she studied air pollution in sensitive alpine environments (tinyurl.com/scienceofsnowmelt), before making her way to the University of Washington to work in the Laura Prugh Lab on wildlife ecology.

Taylor Ganz stopping for a photo during our hike.

Predator-Prey

Now Taylor is deeply entrenched in her research at UW. Taylor’s research is part of a big collaborative project between the Washington Department of Fish and Wildlife and The University of Washington called the Washington Predator-Prey Project (www.predatorpreyproject.weebly.com).

Wolves returned to Washington state in 2008 with just one breeding pair documented. Now, the Washington Department of Fish and Wildlife reports there are at least 145 individual wolves in Washington. The Washington Predator-Prey project was mandated by the state legislature in 2016 with the goal to understand the impacts of the recolonization of the wolves.

Taylor’s part in the project is focused on how deer and elk populations are impacted by predators, like the recolonizing wolves, both directly, by killing them, and indirectly, by influencing their behaviors or other predators.

 “Predators can impact individual prey by killing and eating them,” said Taylor, “…this may or may not impact prey at the population level.” Prey might avoid certain areas or move about differently in the presence of a predators. For example, the presence of a predator may cause deer to spend more time hiding and less time feeding. Of course, this can also have an indirect impact on the deer population by limiting food availability or stressing the animals, which could also decrease deer survival.

Taylor works two 5,000 square-kilometers study sites—one in the more populated northeast part of the state and a second in the less populated north-central part (situated on the east slope of the North Cascades). The northeast site has four wolf packs, while the northcentral site only had only two until recently.

Large-Scale Change

Young trees stood at attention on both sides of the wi­de gravel trail as we walked along. You could still see the stumps from the last harvest and grass-lined the trail.

“I am also interest in large-scale changes in the environment, like fires and timber harvest,” Taylor told me. This is why Taylor suggested visiting an active timber harvest site for our hike.  Though her research sites are in the northeast and northcentral parts of Washington state, much of the sites contain active timber harvest as well.

Areas that have been harvested for timber have the potential to create good habitat and forage for deer and elk populations, at least for a few years. Harvesting timber opens the canopy so that sunlight can reach the ground, which can promote the growth of shrubs and herbs that deer and elk feed on. As the trees grow up, they can provide a place to rest and hide from predators.

The trail as we were getting started on our hike.

Keeping Track of Deer

Eventually, we dropped down into a section of forest with taller, but still not mature trees.  Sword fern and Salal became the dominant understory and I was excited to see a few trilliums still in bloom. Enjoying the shaded trail, I asked Taylor to describe just how she goes about finding deer and tracking them.

“We need to track adult females and younger animals,” explained Taylor.  Males are largely ignored for her study as they are abundant enough to get the females pregnant. Instead, the focus is on capturing and collaring adult female deer and elk and their fawns and calves.

“Elk are primarily captured by aerial darting, and mule deer are mostly collared by aerial rocket netting,” said Taylor. But for white-tailed deer, she often uses ground darting and special traps, called clover traps, to lure deer and capture them. Taylor described them as large boxes with a tripwire that closes the door. Bait, like corn, hay, or something called “sweet feed” is used to attract the deer inside. And once captured, the transmitter sends a text to Taylor, so she knows to check the trap.

When Taylor arrives at the trap, she needs to add a GPS tracker to the deer—this is a priority. To do so the clover trap is flattened and laid down on its side, essentially creating a “deer sandwich.” Taylor is then able to apply an anesthetic before tagging the deer.

The other tool Taylor uses to capture adult deer sounds a bit like a fourth of July firework or something military—”suspended rocket net.” This contraption uses a net with weighted corners that launches during capture overhead an unsuspecting deer.  Bait is used to lure the deer into position below the net before it is launched. “It’s kind of like a bug in a spider web,” said Taylor.

One of several Pacific Trillium seen on the hike.

Tough Captures

Later, Taylor told me about the challenges of trying to capture new animals. First, it takes a lot of detective work, looking for “pellet piles” and tracks.  Then the trap must be “appealing for the deer.” Taylor told me about a time her boyfriend and dad came with her into the field. She had them both dig a shallow ramp in the snow up to one of her traps to make it easier for a deer to enter.

“No, we need a really gentle entry,” she told them as they worked. 

From then on, the trap was referred to as “The Ruby Creek Country Club,” based on its luxury accommodations and location.

Technology Revolution

No matter the method, once captured, the adult deer is collared with a GPS tracker, measurements are taken, and blood and hair samples collected. Vital signs are monitored during the process to ensure no harm. If the vitals deteriorate, the deer will be released. Blood is especially important for tracking pregnancy.

In addition, they will often ultrasound the deer in the field as well. And if pregnant, implant what is called a VIT (Vaginal Implant Transmitter). The VIT talks to the deer’s collar, texting Taylor when the VIT has been expelled during birth. From there, Taylor and her team can often track down the new fawn and add an expandable collar that uses radio telemetry to “speak” to the mom’s collar.

Currently, Taylor has about 200 ungulate “on air.” Each spring they capturing between 30-40 newborn fawns and elk calves.

“Much of the technology is really recent,” said Taylor. In the past, radio telemetry was really the only option for tracking animals, which requires more time in the field with receivers and triangulating positions. Now, GPS collars allow you to know where the animal is every 4 hours, and you can see their positions from the comfort of your own home if you want.

Elk Tracking

Of course, every animal is tracked differently. For Elk, for example, a mix of strategies is used, like aerial telemetry, in addition to GPS collar. It is also helpful to consider the behaviors of the animals for tracking. For example, often elk will move large distances in one day. When this movement stops suddenly in late May, this could mean a calf was born.

On the Look Out

At this point, we had made a few turns on our loop and were getting close to Mima Falls, our designated hiking destination. The forest was still fairly shaded with younger trees which made for a cool walk.

As we hiked, Taylor and I kept a lookout for signs of wildlife, but so far had not seen much if anything. Of course, that does not necessarily mean there is nothing around. I often think about the wildlife that I do not see while I am out hiking. Could there be eyes watching us now as we walk through the forest?

Deer are also on the lookout for other wildlife, especially predators. Of course, they are much more capable of detecting the presence of others than us poor-sensing humans. Interestingly, what a deer detects seems to matter a lot—changing their behaviors in possibly meaningful ways.

A shady section of trail.

Wolves, Cougar, and Bear, Oh My!

Taylor mentioned a study out of Yellowstone that showed that when cougars were around, elk tended to move out into the open, but when wolves were around, they tended to move into more densely wooded areas. This makes a lot of sense, as cougar are ambush predators—stalking their prey and going in for a quick kill. Better to be out in the open where a sneak attack is more difficult. While wolves are coursing predators—chasing their prey to exhaustion. Better to be well hidden where it is difficult to chase.

“We have collared wolves and cougar, as well,” Taylor told me, as part of the Washington Predator-Prey Project.  Using a movement modeling technique called a “step selection function,” Taylor is also looking at how deer and elk are responding to predators. 

“Basically, the model considers a handful of possible routes a deer or elk might take, based on their location at a given time, and compares it to the track they actually take. Then, by looking out how their choices change in the presence of cougar, for example, we can see if they are altering their habitat use.”

She is also looking at if humans are influencing these interactions as well. “One of the thoughts is that they might move more toward people because predators tend to be more human adverse,” Taylor shared. 

Playing Nicely

“Do they play nice?” I asked Taylor, referring to the cougars and wolves.

Though not her primary focus, Taylor told me that one of her collaborators is working to answer this very question. One thought is that cougar might move into higher elevation areas if wolves are nearby, but there are more than wolves and cougar to consider.

“Bobcats, bears, and coyotes too. All of these can eat deer and elk at some point during an ungulate’s life,” explained Taylor.

Previous research indicates that some mesopredators, like bobcat and coyotes, will change their temporal activity when apex predators, like wolves, are around. And that coyotes and bobcats may find human interaction less threatening when apex predators are present.

Fun in the Field

We continued past a couple of gorgeous incense cedar trees growing along the trail, admiring them as we went. The forest had not changed a lot, but the trail leveled off during the last stretch to the waterfall.

I asked Taylor if there were other ways, other than trapping, that data was being gathered on wildlife populations for the Washington Predator-Prey Project.

“There is another PhD candidate, Sarah Bassing, that has over 100 camera traps set up,” Taylor exclaimed. “She is looking at interactions on a large scale.”

Taylor also told me about another way they are using camera traps to track scavenger behavior at carcasses.

Scat is also collected for most of the carnivores in the research project. “There is a couple of key pieces of information it can tell us,” explained Taylor: 1) what did the animal eat, and 2) who left the scat?  You can even use how much scatt is collected to estimate how many individuals of a species are out there.

Overall, “There is going to be three PhD dissertations, one master’s thesis, and additional papers” published around the Predator-Prey project. And hopefully a synthesis paper will be written, Taylor suggested, but probably not for many years.

Waterfalling in Love with Plants

We eventually made it to the Mima Falls. It was not particularly spectacular, but we still took a short break to take in the views and snap a couple of photos, before heading back onto the trail.

View of Mima falls

“Well, this is the type of forest I am used to…but with a little less fern and more shrub” said Taylor as we walked along through the canopy-filtered light. The forest was still mostly well-spaced young Douglas-fir.

“We run habitat surveys in the summer,” said Taylor. “We have documented three-hundred plants in the understory.”

Taylor admitted that after a long fawn and calf capture season in the winter, focusing on plants is a nice change of pace. “It can feel so relaxing,” she remarked, and “I have been able to learn a lot of plants.”

Of course, good sources of vegetation are important to deer and elk success. I asked Taylor if deer have any preference toward certain plants. Taylor mentioned a study conducted by Lisa Shipley, from Washington State University, where captive deer are observed, and what they eat is recorded “bite for bite.”Mule deer tend to prefer deciduous shrubs, such as willows and serviceberry, while white-tailed deer favor forbs such as heartleaf arnica, bunchberry, and even strawberries.

Understory shrubs along the trail.

Migration

As we looped up the trail, we entered a more mature forest with several older, larger trees. While we walked, our conversation migrated back toward a theme, we touched on a bit earlier—migration. More specifically, what are the patterns of migration, and do landscape-level changes impact migratory routes.

In her Okanogan county site, where her focus is on mule deer, Taylor is particularly excited to better understand seasonal migration patterns. “Most animals move up in spring and move back down in winter,” said Taylor, sometimes moving up to 40 miles over the season. This movement is thought to occur due to changes in available browse as spring moves to higher elevations and latitudes—what is known as the “green wave hypothesis.”

One of the larger trees we walked past on the trail.

A Slow Burn

However, with landscape-level changes, such as fire, Taylor suspects movement patterns will change. “I have two ideas about influences of fire,” she said, “both relating to how they reduce canopy cover. One way is it could make really good forage and they may be attracted to these areas.”

On the other hand, “the canopy cover can really block snow from accumulating.” With reduced canopy cover post-fire, snow may be deeper and reduces food access. Deep snow also reduces deer mobility when they are trying to move fast to escape a predator. The fact that “deep fluffy snow favors predators over prey is well established,” remarked Taylor. She hopes to parse out these effects.

Down Logs

Similarly, down logs can also be thought of as both helpful and harmful to deer. We were noticing a lot of down logs along this section of the trail.  First, logs “can alter the way they are moving,” hindering their ability to escape a predator. But second, logs recycle nutrients to the forest and provide habitat for other forest life.

Lots of down logs in the mature forest.

Why should we care

By now we were more than an hour into our adventure, and I realized I had not asked one of the important questions—why should we care? And, of course, Taylor was quick with a response.  She really knows her stuff!

“There are a couple of reasons,” she began. “Their role in the ecological community and also people really like deer and elk.”

“They are really cool because they are a species people can see and relate to and understand in a way a bobcat perhaps is not,” said Taylor. “I can go out in the field and see 50 deer!”

As for ecological benefits, deer and elk are “middlemen” in the food web—affecting both the landscape of vegetation while also supporting predator populations. There is also some evidence that they are important to transporting seeds and moving nutrients around.

Deaths

One of the most basic concepts of wildlife populations is understanding death.  So, it is not surprising that throughout our hike together, Taylor and I discussed deer and elk death at length.

“Are there any threats to deer and elk?” I asked at one point.

“There are a number of different diseases that regulate a population,” Taylor offered. Though not found in Washington State, Chronic Wasting Disease (CWD) is a potential concern for many populations of ungulates, as it spreads through saliva and tends to be more of a problem in dense areas. 

This is, of course, where predators may come in. By potentially keeping deer and elk populations less dense, diseases, like CWD, may become less of a threat. In addition, problems with too many deer, like wildlife conflict, eating crops, and vehicle collision are also lessened. There is some evidence, especially in the Eastern United States, that predator return could save lives and money.

Necropsy

Part of Taylor’s research is looking at causes of death in deer and elk.  To get a handle on this, she has also been doing necropsies—full field dissections—on the bodies of dead deer and elk found in the field. When Taylor told me about her involvement in this, I swear, her eyes lit up. “I really enjoy it,” she remarked.

The main goal of the necropsy is to see how the deer or elk died. If the animal died from predation, there will be lethal bite marks present, if that part of the carcass hasn’t been eaten. If it died from starvation or another stressor, the bone marrow turns from a “candle wax like material into red jelly.”

Another cool side project is to sample lethal bite marks for DNA and use it to identify the individual or species that killed the animal. 

When I asked Taylor if she found any pattern in causes of death from the necropsy, she could not say anything conclusive, as the final stages of data collection were still underway. However, she did note that car collisions were up there, along with predation, and occasional evidence for starvation or disease. How much each of these limits deer and elk populations is still being investigated. “We are seeing some evidence that there is some nutrient limitation,” said Taylor, “but not exclusively.”

Overall, deer and elk are not very threatened. White-tailed deer have, in fact, expanded their range. Mule deer have been on the decline, but they are not considered imperiled, at least not yet. 

Good Habitat

We hiked on, meandering through an interesting section of regenerating forest next to a thin stand of tall trees. I asked Taylor if the area would be suitable for deer and elk and was met with a resounding “yes.”

“This is a good place for cover and lots of good food to get into,” said Taylor. Also, “they love edge habitat.”

However, “in another 15 years it may not be great,” explained Taylor. The trees were planted at a high density, so as they grow taller and without thinning, they will likely shade out the understory, creating a food desert for the deer. 

Toward the end of our hike, we walked through a section of dense forest, like what we might expect for the stand of trees in front of us. The forest floor was dark and there was little vegetation—not a good habitat for deer.

Section of regenerating forest along the trail.

Births

It was during our walk through this dense dark forest, after passing a wet swampy area that white-tailed deer would most assuredly love, that I inquired about the other half of the population puzzle—births.

In biology class, you are taught that there are K-selected species and r-selected species. K-selected species have few offspring, a low growth rate, and stabilize around a carrying capacity. While r-selected species have many offspring, a growth rate, and their population size tends to fluctuate more widely.

Where do deer and elk fit into this picture? Somewhere in between. “They aren’t laying thousands of eggs,” said Taylor, but, like an r-selected species, white-tailed deer are very fecund.  On average a white-tailed deer will give birth to 1.6 fawns a year. Their maturation time is also impressive as white-tailed deer in good condition can breed at 6 months old. Mule deer are similar with a longer maturation time, but similar growth rate. Elk, on the other hand, do not twin and may not have a calf every year—putting them on the more K-selected side of the spectrum.

White-tailed deer also fall more to the r-selected side of the spectrum when it comes to recovery. “They are moving further westward, and their range is expanding,” reminded Taylor. Some people even compare them to rats or rabbits in their ability to reproduce. “I don’t know to the extent they are overshooting the carrying capacity of the landscape, but the potential is there,” said Taylor. 

The Sign

Over two hours into our discussion and we were nearing the trailhead. It was so much fun talking to Taylor and listening to her speak with such focus and passion, I could not believe we were near the end. And we had not seen a single deer, let alone any deer track or scat!

However, just as we were wrapping up, Taylor found something—hair! Was it a sign of wildlife? Not exactly, but it did illustrate a cool feature of deer and elk.

“So, this is probably just domestic animal,” Taylor explained, but “one thing you can do to tell if it is predatory or prey is to fold it.” Predators have several layers of hair to protect against the elements and keep them warm. “These hairs won’t kink very easily,” said Taylor. “Deer hair,” on the other hand, “are really slippery and hollow, so they kink easily.” Deer hair will also fall out easily if a predator grabs them.

And there you have it. Only a few minutes later and we were back at our vehicles saying our goodbyes.

Predator fur discovered on the trail.

Complexity

Predator-prey relationships on the surface seem so simple. One species pitted against another; it can seem like an obvious win-lose situation. But speaking with Taylor, this kind of thinking dissolves—it is much more complicated than that.

Predator-Prey relationships are more like a dance, perhaps a tango, but with more than one partner. What happens to one population affects another, both directly and indirectly, which, in turn, may affect something else altogether. Thus, something as simple as the amount of snow on the ground or the density of trees in a forest has the potential to create a ripple effect over the entire ecosystem. As the old adage goes: we’re all in this together.

Taylor Ganz is a Ph.D. candidate at the University of Washington in Wildlife Science. She has a Masters in Environmental Science from Yale School of Forestry & Environmental Studied,  a B.S. in Mechanical Engineering from the University of Southern California, and a B.A. in Physics from Lewis and Clark College. She also has multiple years of experience as a Senior Field Instructor of National Outdoor Leadership School, and still enjoys teaching for them on occasion.


Science in the Sonoran Desert

I am drawn to wild places, like a moth to a flame. Like so many, my heart beats for the beautiful landscapes of my home in Oregon. The mountains call out to me; the forests whisper their welcome; the rivers run deep in my heart. I have spent a lot of time hiking in the Pacific Northwest, sweating it out on the trails. 

While I love the lushness of the Pacific Northwest, I am also drawn to alien landscapes. Visiting places that are new to me awakens my mind and body to new sensations and curiosities. 

A Foreign Land 

The Sonoran desert of Southern Arizona is about as foreign to western Oregon as it gets. So, for a couple of weeks in March, I traded my muddy forest trails filled with towering conifers for the dusty desert of the Southwest in hopes of discovering a new place to love. 

The sun hung low in the sky, when I met up with environmental scientist and Parsons Field Institute Coordinator, Mary Fastiggi and her colleague, Director of Development and Marketing, Adele Dietrich, for a hike in the McDowell Sonoran Preserve in Scottsdale, AZ.  Our plan was to hike a small three-mile loop starting on the turpentine trail, and along the way, for Mary to show Adele and me a bit of the scientific work the Conservancy was invested in. 

The Hike

  • Trailhead: Granite Mountain Trailhead, McDowell Sonoran Preserve, Scottsdale, AZ
  • Distance: 3 miles+
  • Elevation Gain: 200 to 400 feet
  • Details: There is ample parking at the trailhead and restroom facilities at the trailhead, but no water available. Educational signs and placards are available to read and learn more about the preserve. The preserve is open from sunrise to sunset. Automatic gates close off the preserve at this time.

Orientation

Before we set off, Mary oriented me to the Conservancy as a whole. “We are a stewardship organization that works in the Preserve,” explained Mary; the land itself is owned by the City of Scottsdale.

Started by local citizens, passionate about the land and concerned about land development; over the last 30 years, the McDowell Sonoran Conservancy has advocated for the care of the preserve, with the City of Scottsdale now owning roughly 30,500 acres of land in the Sonoran desert. That is ⅓ of the land area of the City of Scottsdale, and the equivalent of about 36 Central Parks.

Though the original focus of the McDowell Sonoran Conservancy was land preservation, its mission has “shifted from advocating for  land preservation and then expanded to education and science.” Here is where Mary’s work really comes in. She is part of the science pillar of Stewardship at the Conservancy, with a goal of conducting ecological research in order to inform and inspire a deeper understanding of the Sonoran desert.

Urbanization

And there is a lot to be learned! One of the biggest threats to the Preserve and also a priority of the Parsons Field Institute, where the Science pillar is housed, are the impacts of urban stressors. The Phoenix area is the second-fastest-growing area in the U.S. so this is a major focus for the Institute. 

Mary pointed to a map of the preserve found at the trailhead: “This is the Preserve in green,” she said. Though there was a lot of green on the map, you could also see that there was also a lot of land that was not protected, and thus open for development. 

Trailhead Map

Connectivity

“Wildlife corridors are really really important for wildlife,” remarked Mary. With the Preserve adjacent to the Tonto National Forest and other protected land, the function of the Preserve for creating connectivity is paramount. “We want to make sure that connectivity stays connected,” expressed Mary. 

With that in mind, there has been ongoing research on the use of the Preserve as a wildlife corridor and the impacts of nearby urbanization on wildlife.  One break in the connectivity that is a concern is a road, Rio Verde Drive, that cuts across the Preserve.”

“We have done a Mule Deer telemetry study,” said Mary, to see how the road impacted migration. The study revealed different patterns of behavior between the sexes of the deer when it came to crossing the road, confirming that the road does pose a challenge. 

However, one of the projects Mary was most excited about was a new Sonoran Desert Tortoise study, funded by the Arizona Game and Fish Department’s Heritage Grant.  Mary directed my attention out past the map toward an open expanse of land. “It all looks like a desert,” she said, “but all this land can be developed and will be.”

The plan is to use telemetry to track the tortoises over time. By following the movement of tortoises, the Conservancy hopes to gain a better understanding of the impact of urbanization and other pressures on desert life.

Native Plants

After my orientation, Mary, Adele, and I all decided to head out on the trail. Enthusiasm was high as we began our walk along a dusty desert trail heading west. 

However, we hadn’t made it more than a few yards when a colorful red color caught our eyes. According to Mary, it was a Hummingbird or Chuparosa plant!  The Chuparosa is a beauty! Not just to me, but to hummingbirds, as well, who are attracted to the vibrant red blossoms of the plant. It is also “promoted for yards,” said Mary. 

Surrounding the Chuparosa was a lot of golden-colored dry grass. “Red Brome,” stated Mary, “does pretty well and is naturalized across the Preserve.” This non-native grass was one of many we saw through the Preserve during our hike.

Tens of feet beyond the Chuparosa and grasses stood a shrubby plant. Mary grabbed a few leaves and held them to her nose and inhaled deeply. Invited to do the same, a sweet earthy, smell filled my senses. It was a Creosote bush. Apparently, the smell emanates from a waxy coating that protects the leaves from losing moisture during drought. 

Interestingly, after a rainfall, the desert takes on the “fresh smell” of creosote, explained Mary. A reality I can corroborate, as the next week, following a rainy night, I went backpacking in the Superstition Wilderness and was enraptured by the earthy scent of what I could only assume was the creosote.  It smelt like a fresh start.  

Mary with a Creosote shrub.

Different Communities

We also saw some Yucca. Though not in bloom during our hike, in the spring, the Yucca grows clusters of flowers on stocks up to 6 feet high. These flowers are pollinated by the yucca moth which also depends on the yucca fruit seeds during their larval stage—an interesting mutualistic interaction. 

We also saw the narrow-leaved turpentine bush. Though also not in flower, its leaves gave off a sweet smell, which is also an attractant for birds and insects that use the bush for food and shelter. Contrarily, it also attracts botanizing humans as well.

“You won’t find as much of these in the southern part of the Preserve,” Mary stated, referring to the Yucca.

“You don’t really see this one a lot lower either,” Mary pointed out, referring to the Turpentine bush. 

Within the preserve elevation changes, as well as differences in rainfall and temperature, have resulted in many different community types or subtypes. “There is about a 700-foot difference in elevation from the lower parts of the Preserve,” explained Mary.  Thus, depending on where you are in the Preserve, you will see a different mix of plants and animals. 

Yucca found along the trail.
Turpentine bush is common in the preserve at higher elevation.

Defining a Desert

So with all these differences, what exactly defines the Sonoran Desert? And why should I care?  

At this point, we were nearing one of Mary’s Field Sites, but before I could focus my mind on what I was about to see, I just had to know—Why the Sonoran?

Mary stopped to answer my question: “The Sonoran desert is one of the most biodiverse deserts in the world,” said Mary.  “It is also the only desert with saguaro cactus,” an indicator species, found consistently across the Sonoran desert. 

“It is fascinating,” Mary went on, “it gets quite a bit of rain in some areas compared to other deserts.” It also has two rainy seasons—the winter rains and the monsoon rains. These rains bring about changes in the vegetation. 

“Rainfall is very localized,” said Mary, contributing to habitat biodiversity, and ultimately, species diversity. Annually, you might get only 3 inches in one location and 20 inches in another.

Thus, though some things are consistent across the Sonoran Desert, like the Saguaro, it is the differences across its range that make the Sonoran Desert a spectacularly special place. 

Desert Crust

Satisfied for the moment, Mary brought my attention to a checkerboard pattern of nail heads sticking up out of the soil. Behold—the soil crust experiment. 

“Soil crust, which is a community of living things that are very very small and usually slow-growing, covers a lot of the Sonoran Desert,” Mary shared as we looked out on the microcosm. “In an untouched area you will find a lot of soil crust, but with human impact it is one of the first things eliminated.”

Despite the smallness of the community of microbes that create the soil crust, loss of soil crust is a huge deal in the Sonoran Desert!  Not only is soil crust incredibly slow-growing (once it is gone it is gone), but soil crust retains water, aids in nutrient cycling, and stabilizes the desert soil. 

“Have you heard of the Haboobs?” Mary asked.  I hadn’t. “The haboobs are giant clouds of dust— a wall of dust—and they take over the city.” They literally “stop traffic” Adele offered. Why do these dust storms occur? Lack of soil crust.  

So again, soil crust is small stuff, but kinda a big deal. As Mary puts it: “soil crust is the skin on the desert.” 

Growing Soil Crust

In any event, the “nailhead” soil crust study is an attempt to better understand how soil crusts form. “To see if we can make soil crust faster,” said Mary. 

Basically, the Conservancy, with the aid of volunteers, and in partnership with Northern Arizona University and the City of Scottsdale, collected soil crust from a city excavation site. Then took that material and treated it in a variety of ways. And now are watching to see how the microbiota of the soil crust will respond. Some examples of treatments include:1) seeded the soil crust with native seeds, 2) soil crust rolled out on burlap material with minimal disturbance, and 3) crushed and spread soil crust.

We got up as close as possible to see if we could see any developed soil crust but were unsuccessful. Normally the soil crust in the Sonoran Desert is thick and black with microbial growth. At this stage in the study, there was still not a lot to see. 

Mary also mentioned that other scientists are trying to characterize the biota of soil crust in various deserts around the country. There is still a lot we don’t know about soil crust communities and how interchangeable, if at all, they are from region to region. 

Site of desert crust experiment.

Restore a Desert

We continued on past the sparkling nail-heads of the soil crust experiment, tracing the path of power lines overhead.  We even passed a pool of water along the trail, an unusual site to see in the desert. 

Eventually, we made it to the next destination on our Preserve tour—the RestoreNet plots. This was the last major site Mary planned to show Adele and me. The RestoreNet site is an old camping and recreation area from before it was part of the Preserve. It was chosen from 67 sites identified via satellite imagery because it was impacted. 

Mary explained the purpose of the experiment.  “A lot of restoration projects are very localized. Most don’t go very well. There is a lot of seeding failure.” RestoreNet is an answer to this problem. By involving a “network of sites” across the country, the project hopes to provide land managers with more comprehensive information regarding restoration site treatments.

RestoreNet experimental site

Treatments

“We are in charge of four RestoreNet sites in the Sonoran Desert,” said Mary.  Each is different in terms of type and degree of impact, but each received the same treatments.  At each site, “we have 36 different plots,” said Mary. Half are seeded with a warm mix and half a cool mix, with the exception of four untreated plots.

Mary took us around the site showing off the different treatment plots. There were mulched plots, plots with connectivity modifiers (basically metal stood up on end), seed-only plots, and pit plots. 

Though it is still too early to tell what the RestoreNet results will ultimately suggest, there is preliminary data that is interesting. The first year of data gathered on the Colorado Plateau, for example, showed the use of pits did the best.

Though there wasn’t a lot of growth at the RestoreNet site we visited, we did see a few different lupine species in one plot, one of which is part of the cool species seed mix (Lupine sparsiflora or Coulter’s lupine). Mary said she would return in a week to conduct the spring plant monitoring, collecting data at all four sites.

RestoreNet Pit treatment

Dry Winter

With an overall drier winter, the lack of plant emergence in the RestorNet site wasn’t necessarily surprising. The plots are not watered, so just like in a normal situation, the plants are relying on the weather for water.

Thus in addition to the plots, a rain gauge is set up on-site. The rain gauge was a simple plastic gauge holding an inch of water in the center with the ability to collect water beyond this capacity.  Adele took a measurement that day and recorded it for the site: “0.57 inches.”  “We have had rain here,” said Mary pointedly, so hopefully in a week, the plots will be in bloom. 

Adele measuring precipitation using a rain gauge

Desert “Old Growth” 

After visiting the RestoreNet site, Mary, Adele, and I took a right onto a trail heading toward Granite Mountain. The path was lined with cactuses, including Buckhorn Cholla and, of course, Saguaro. 

I was especially enchanted by the Saguaro cactuses that stood proudly around us. There were so many of different sizes and shapes; many had several arms, the older Saguaro had cavities, and a few were in various states of decay. It reminded me a bit of the old-growth forests of Oregon and I asked Mary to talk more about the significance of Saguaro to the Sonoran Desert ecosystem. 

Mary next to a many armed saguaro

Habitat

Saguaros provide excellent habitat for a variety of animals: from Gilded Flicker and Gila Woodpecker who build their homes in its flesh, to Owls and Martins who take advantage of empty cavities. Bats consume the pollen and nectar of the saguaro blossoms and many species eat the moisture-rich fruits of Saguaro in the summer months. A decaying saguaro reveals its internal woody skeleton; this becomes home for other desert species, including insects, snakes, and rodents. 

When other sources become scarce, home animals rely on Saguaro for a freshwater source. The Saguaro’s pleated stems allow for expansion and contraction depending on the state of hydration. By feeding off the flesh of the Saguaro, animals like pack rats and mule deer take advantage of the Saguaro’s adaptation to drought. 

Saguaro Life

The life of a Saguaro is long and requires certain conditions for success. According to Mary, Saguaros rely on nurse plants for successful germination. Without this protection, their chances of survival are slim.  

The Saguaro also takes a long time to grow to their natural height of 35-40 feet over their up to 200-year life spans. Signs posted at the park state that “At age 40, a saguaro is only 3 feet tall.” Arms don’t appear until around 75 to 80 years according to Mary. It is no wonder that Saguaro is a protected species in Arizona.

Young saguaro surviving next to a nurse plant

Notice and Wonder

As we continued up the trail, Mary, Adele, and I continued to observe and identify the plants around us. I asked what else we might notice and wonder about in the Sonoran desert. 

“This is the green time of year,” said Mary. “It can be greener, but it is greener than it has been.”

Adele pointed out the small size of the flowers and leaves in the desert. One key way to appreciate the Sonoran Desert is “taking time to see the variety in what others might go by,” Adele suggested. You have to stop and take notice. 

So that is what we did. 

We saw fairy dusters with their pink and red blooms; Palo Verde trees with their bright green stems and mistletoe growing in their branches. We saw baby Saguaro and old saguaro pods littering the desert floor; ocotillo with their stiff, spindly branches fanning up toward the sky; and many other unidentified blossoms on the desert floor.

Chuparosa brings color to the desert with it’s bright red blooms.

Burning Questions

As we continued our way along, making our loopback toward the parking lot, the conversation shifted back to the priorities of Parsons Field Institute. It was at this point that we noticed several large plots of non-native grasses along the trail. 

“Typically in a non-infested area, you would have a lot of space between plants,” explained Mary, but “invasive grasses take over those spaces.” 

However, this is only the beginning of the problem. Besides outcompeting many other native plants, invasive grasses “turn brown and catch fire.” 

“The Sonoran Desert ecosystem is “not adapted to burn at that intensity and it allows the fire to spread,” said Mary. She added: “The problem is that it is a positive feedback loop,” Essentially, this means that fires promote more fires, as plants adapted to fire like the non-native grasses return following a fire in full force, while others, like Saguaro, take a long time to recover. The result: many areas of the Sonoran desert are shifting toward grassland habitat. 

This is “a massive threat to the Sonoran Desert,” Mary stated grimly. Managing the problem is a challenge as well. With a seed bank already established in the park, removal by volunteers is possible, but a constant battle. 

Invasive grasses cover the soil.

Climate Change

Shifts in climate change is another area of focus for Parsons Field Institute, but it is hard to say what the long-term impacts of climate change will be. The main focus at the moment is monitoring in order to gather long-term data sets.  With the help of over 500 volunteers, Parson’s has been able to monitor butterflies, bats, birds, arthropods, and plant phenology. 

Stewards

The sun had sunk lower in the sky, and the large granite boulders that lined the trail were casting long shadows, as we neared the end of our hike.  

However, before parting ways, I asked Mary and Adele to tell me more about volunteer opportunities at the conservancy.

According to Mary and Adele, people are involved in the Conservancy in many different ways. Some Stewards learn about the park through educational experiences. Other Stewards participate in research or volunteer restoration hours. Whatever their involvement, the goal is still the same—connection.

Sometimes the connection is deep. Adele told me of instances where people whose experience as a Steward has “shaped their retirement.” While Mary spoke of Stewards that were so deeply involved in science projects, that they consequently were listed as authors on publications.

A big part of the Steward program is “getting people to have a relationship with the land,” said Adele, and there are a lot of different ways that can happen.

No Place Like Home

After parting ways with Mary and Adele, I wandered the trails of the Preserve for a bit longer on my own—opening up my eyes, ears, and nose to the sights, sounds, and smells of the desert. I pondered the outstretched arms of the saguaro and marveled at the rounded rocks that lay jumbled across the landscape. 

I spent several more days exploring the Sonoran Desert acquainting myself with its various moods and rhythms. 

Adele had mentioned during our hike how my blog post was providing “an outsider perspective.” I can appreciate that. 

But, at the same time, when I think about the bold desert sunsets and the moonlit nights, or of early morning bird song and the afternoon drying heat—I don’t feel like an outsider. I feel like I belong —in the wild Sonoran Desert.

Mary Fastiggi is the Parsons Field Institute Coordinator for the McDowell Sonoran Preserve. She has a B.A. in Environmental Studies and History from the University of Michigan and an M.S. from Arizona State University in Sustainability.

Adele Dietrich holds an MBA from Alaska Pacific University, is a Certified Fundraising Executive (CFRE) and holds the Chartered Advisor in Philanthropy (CAP) designation. Her goal is to connect individuals with their philanthropic passion.

Hike Back in Time at Petrified Forest National Park

Scenery while hiking cross-country through the Flattops.

For me, trying to understand geological time is a bit like trying to fit a square peg into a round hole. It takes some serious reshaping before the pieces start to fall into place.

When I met up with Dorenda and Matt Walters, my hiking guides at Petrified Forest National Park, little did I know, just how much mental craftwork I was in for—225 million years’ worth! That is how much life history exists in the park—a seriously mind-boggling sum.

A Long, Long, Long Time Ago

Before setting out on our hike, Dorenda and Matt Walters arranged for a tour of the park’s museum collections. Matt Smith, curator and paleontologist, led us on this venture. 

To start, Matt Smith shared a “mental-gymnastic” he uses to try and get his mind around the 225-million-year history:

“T. rex died 66 million years ago,” he explained. “His oldest cousin lived during the Triassic in (what is now) the Petrified Forest 220-225 million years ago. We are looking at more time between T. rex and the oldest dinosaur and T. rex and us. T. rex is closer to the iPad than its earliest ancestor.”

This is the timeframe we are working with—almost four times the amount of time it takes to go from dinosaurs to humans. So, as you can imagine, back then, the Earth was a completely different place.

“This planet was on the other side of the galaxy,” described Matt Smith. The continents were united into one supercontinent—Pangea.  It was the dawn of the dinosaurs. Mammals were just getting started. And flowering plants had not even shown up yet. 

In other words, it was a long, long, long time ago.  

The Box

After his brief introduction to time, Matt Smith led us into the Museum Demonstration Lab, or “the box,” as he called it—a small white room with windows and desks facing outside that allows visitors to glimpse in the “behind the scenes work” paleontologists do at Petrified Forest. 

“Fossils are our jam,” said Matt Smith, before introducing us to the room’s current occupants—a metoposaur skull and a phytosaur skull. Each sat on separate desks facing the window, cradled inside their plaster jackets.

Metoposaur

The metoposaur’s fossil skull was roughly triangular and flat or, as Matt Smith put it, “shaped like a toilet seat.”

Overall, metoposaurs were large amphibians, “up to 10 feet” in length, with rough textured skin similar to the bone underneath, and a body plan like a modern-day crocodile, only stouter.

As carnivorous feeders, metoposaurs would sit on river bottoms, “open up their mouths like a bass and feed off whatever came into their mouth,” said Matt Smith.

He pointed to a deeper trench hidden in the texturing of the fossil. He explained how this trench was part of a lateral line system, like fish have. This system would have allowed metoposaurs to sense their prey, even in the murkiest of waters by detecting changes in pressure or electrical pulses.

Metoposaurs were “common everywhere up until the end of the Triassic,” said Matt Smith

Matt Smith standing next to the metoposaur skull in “the box.”

Phytosaur

The phytosaur skull had an even more unusual shape. It looked a bit like an alligator but with a very long snout, and nostrils toward the back of the head, instead of the front.

Phytosaurs were huge, maybe “25-30 feet long,” with long tails and sharp teeth; again, with the body plan of a crocodile. “They were fish eating specialists,” said Matt Smith, “Crocodile-like 80 million years before crocs.”

He went on, “They don’t have common ancestry (with crocodiles) … these guys turned ‘crocodile’ by stretching out their premaxilla.” Crocodiles, on the other hand, stretch out everything in the snout. That is why a phytosaur has nostrils at the back of the head and crocodiles the front.

Crocodiles and phytosaurs are an example of convergent evolution—similar environments, resulting in similar structures on totally separate locations and timelines.  When a body plan works, it works!

The phytosaurs are one of several archosaurs that are found at Petrified Forest. Phytosaurs are not dinosaurs and exist on a separate branch of the archosaur family tree. They are a group of reptiles that includes dinosaurs as well as modern birds and crocodiles. Phytosaurs are one of many Triassic archosaurs found in the park. The only two living archosaur groups are crocodilians and birds.

Matt Smith with the phytosaur skull in “the box.”

A Curved Femur

After our visit to “the box,” Matt Smith brought us into the collection rooms. Lined with metal cabinets, the collection room contains hundreds of catalogued artifacts and specimens found in the park.

The first set of specimens Matt Smith introduced us to were fossils from an azendohsaurid reptile.

“This animal wasn’t known in North America,” Matt Smith explained, until 2014 when a weird vertebra, discovered in a loan return, piqued the interest of park staff. Before long, a fossil site filled with azendohsaur fossils was discovered, and 40 different field jackets with specimens were collected.

Now, all these specimens stood in front of us—organized and packed into a short metal cabinet with wheels.  Matt Smith pulled open the first drawer. Dozens of tiny femora (upper leg bones), broken from the weight of time, lay arranged in small, labeled boxes.

Matt Smith pointed to one of these fossils with a slight bend in it. “This curve is due to natural disease like rickets,” he remarked.

Other drawers contained other parts. All put together,the azendohsaur was about the size of a medium dog “with a long neck and sprawling leg posture,” described Matt Smith.

Azendohsaurid reptile fossils in a drawer.
The curved femur.

Modern Dinosaurs

Next Matt Smith directed our attention to a much larger metal cabinet. “Birds. We have a lot of birds,” he exclaimed as he opened the cabinet and pulled out a drawer. And there they were lined up in a row—dozens of taxidermy birds, from the Northern Flicker to Common Ravens. Matt Smith picked up one of the specimens, a Saw-whet owl—a bird never-before-seen at the park. That is until it was found recently on park grounds, having died of unknown causes.

Now you might be wondering, “why save a bunch of dead birds anyway?”

Well, museum collections are like information investment accounts. The value of the specimens when first catalogued might seem small, but over time, with changes in technology and new scientific questions, a greater value is realized.

As Matt Smith put it, pointing to the tray of birds, “Hopefully, these will help answer questions in the future.”

In addition to modern-day bird specimens, a 220-million-year-old dinosaur fossil was found in the park—the ancestor to modern-day birds. “We have had dinosaurs here longer than anywhere else in North America,” Matt Smith stated. “And we have proof.”

Drawer of varied bird specimens.

People

Closing the bird specimen cabinet, Matt Smith directed us to another similar non-descript case.

Inside was a collection of pottery arranged carefully on pull out trays.  The vessels were a variety of shapes and colors, each one carefully decorated.

“People have also been walking around the forest for a long time,” said Matt Smiht. Though not as staggering as the dynasty of archosaur life, human history in the park goes back 13,000 years. 

And they are still around today. There are “37 tribes on the land,” he states, referring to the number of tribes that are affiliated with the parklands.

Among the artifacts in the collection were examples of Adamana Brown-style pottery, a form of pottery dating back to around 250 BCE. These early brown and gray ceramic pieces date back to a time when pit house villages sprung up and seasonal farming was a focus.

Later, from 650-950 CE the ceramics changed from plain brown and gray to decorative black-on-white designs and corrugated pieces, a style associated with pueblo development. Then from 950 to 1300 CE ceramics diversified even more, with black-on-red and polychrome examples showing up in the Petrified Forest archeological records.

Many of these forms stood on attention against the stark gray industrial cabinetry.

Drawer of varied pottery.

A Legacy

Matt Smith pointed out a piece that was yellow and brown—a Hopi-style ceramic.  “This one is probably 400 to 500 years old,” He said, “fired at a slightly higher temperature,” than the black and white pieces.

He went on to explain how this style of pottery was almost completely lost. He pointed to another piece from the 1960s by Fannie Nampeyo—“the last one who knew how to pot in this style.” Fanny Nampeyo learned from her mother before her, also called Nampeyo, who revitalized the ancient Hopi style in the 1890s. Without the Nampeyo legacy it is possible the Hopi pottery tradition would have been lost.

Turkey Feet or Lung Fish

Before Matt Smith shut the cabinet, a small piece of corrugated pottery caught my eye. Decorated with simple lines that resembled chicken feet, I asked Matt Smith to tell me more about it.

“This is cool to me,” enthused Matt Smith referring to the markings. “One archeologist thought they were turkey feet” he said, but the number of talons does not add up.

“It is not a common design element,” Matt Smith said, while he doubled back to another cabinet behind us and began rummaging around, so “I have my own theory.” He pulled out a small fossil that looked a bit like webbed feet— “lungfish teeth,” Matt Smith exclaimed. He went onto explain how lungfish teeth are common Triassic fossils found in the park and have often been found associated with prehistoric structures.

Could these mystery markings be paying homage to lungfish teeth fossils? We just don’t know.  

Lungfish teeth fossils.

Type Cabinet

We had been at it about 30 minutes, when Matt Smith took us to the creme de la crème of the museum collection—”the type cabinet.”

“In natural science, you have got holotypes, explained Matt Smith. “They are the sample—skin, skeleton, genetic material, fossil—that was used to describe a new species. They are the archetype of that animal.” Every other specimen found must be measured against existing holotypes in order to determine if a species is new or not.

Holotypes

Matt Smith showed us a few of the holotypes housed in their museum collections. “some are pretty miserable, said Matt Smith, “a single tooth or claw” might define an entire species. One holotype that Matt showed us was Vancleavea campi, a species of reptile that may have lived more than 11 million years during the Triassic. “Covered with armor… it isn’t related to anything alive today…” said Matt Smith—it was essentially “bulletproof.”

Of course, the challenge with modern-day holotypes is often ethically obtaining a specimen in the first place, especially when the species is rare. To get around this, people often must be creative and very patient. For example, Matt Smith told me about how scientists found a new species of iguana on the Galapagos Islands about 10 years ago. In order to obtain the holotype, they had to find a living iguana that they felt would work, put an RFID chip in it, and sit back and wait for nature to take its course. I believe they are still waiting to this day.

A drawer of holotypes, including Vancleavea campi.

Mussels

Matt Smith also showed us a diversity of Triassic freshwater mussels holotypes. Modern freshwater mussels are “more diverse than anywhere else in the world in North America,” said Matt Smith. But they are in trouble. “These are going extinct faster than any other group of animals in the U.S.,” Matt Smith explained, despite the fact that they are “evil geniuses” according to Matt Smith, able to disperse their young by smuggling a ride on migrating fish.

Collection of mussel holotypes

Plants

Matt Smith also showed us some plant fossils, among them fossil trees. There are “14 species of trees in the park,” according to Matt Smith. However, despite their abundance working with plant fossils is difficult. “Plants never die in one place… they die in parts,” said Matt Smith. You aren’t going to find a complete plant body like you might for an animal. Thus, a plant holotype requires some closer examination. The tree holotypes Matt Smith showed us during our tour were thin sectioned specimens, in order to see the grain of the wood.

The type cabinet looked pretty similar to others we had seen with the exception that it was on wheels. Why? In the case of an emergency, wheels provide a quick getaway. “77 species would be lost if we didn’t have this,” stated Matt Smith.

Drawer of plant holotypes.

What’s the Point?

And on that note, Matt Smith took us around the corner to the back of the collection room we were in. We walked past some furniture built by the CCC (Civilian Conservation Corps) in the 1930s—another layer of human history at the park—and over to a final cabinet filled with artifacts.

The final cabinet we visited that day was filled with small clear packages of artifacts, each filed in equally small boxes. Shells from the Gulf of Mexico, pipestone from Wyoming, obsidian from Flagstaff, turquoise from New Mexico and, of course, petrified wood—each artifact shaped by human hands– telling the story of human migration and technological change in the area.

Matt Smith pulled out several points and talked about their various uses. Like the pottery, Petrified Forest National Park hold a record of points/tools dating back 13,000 years from “Clovis through Folsom, basket maker, and Puebloan.”

Matt Smith described a place in the park, a playa, where some points and a lot of chunks of material (lithic scatter) have been found. “There are petrified wood deposits… and a little rise,” said Matt Smith.  The playa would have been filled with water 13,000 years ago, so it would have been the perfect place to both hunt and make points for hunting.

One of many points Matt Smith showed the group.

The Lab

The final stop on our whirl-wind tour was the paleontology lab, so we stepped outside and made our way across the park campus. Before long, Matt Smith ushered us into another non-descript building.

“So, this is the Prep Lab,” Matt Smith exclaimed. “Most of what we do in here is paleontology… we do basic conservation work for non-paleontological stuff… but we do the whole shebang for fossils–from the grave to the cradle.”

Gumby and Reynaldo

Looking around the room, it looked a lot like any other well-lit lab space, but with a couple rather large fossils sitting out in their plaster casts on lab benches.

 “This is Gumby, a phytosaur skull,” said Matt Smith. The fossil was in disrepair—the back end of Gumby lay in a plaster jacket in two pieces. Apparently, Gumby got its name because it is bendy, but also likes to break; so, after two or three breaks, the staff decided to create a mold of it. Matt Smith told me that the plan is to use casts from Gumby and several other individuals to create a replica of a phytosaur skull for display. He showed me a partial cast of a phytosaur jaw made from two fossils cobbled together.

Matt with a cast of a phytosaur fossil.

“And here is Reynaldo,” said Matt Smith, “he’s a big sexy beast.” Collected in 2016, Reynaldo has been an on again off again project for several years now— “probably three or four hundred hours” put into preparing the fossil, said Matt Smith but now “it is really close.” With a little more reconstruction of the face—and lots of glue and plastic—the staff hopes to get him stabilized soon.

The reconstruction of Reynaldo.

The Small Stuff

As Matt Smith grumbled a bit about the frustrating nature of larger pieces like Reynaldo and Gumby, he directed the group toward the back corner and another shiny metal cabinet.

“My heart lies more with stuff like this,” said Matt Smith as he pulled out a drawer and pulled out a small vial with a tiny fossil inside. “All these tiny little fossils to me are a lot more fascinating…

I can prep them out in a day or two, and I can store a lot of them, and it’s just way more rewarding if you ask me,” he explained.

A drawer of small fossils.

Origin of Lizards

In order to study the small fossils, they are sent to another lab for a Micro-CT. This sort of imaging is like a regular CT scan, only more intense and the scan machine is much smaller, fitting on a desktop. The Micro-CT can get finer detail with micron-size slices of images captured. The information from the Micro-CT can then be used to print a blown-up plastic version of the fossil using a 3D printer.

Matt Smith held up an example of a 3D printed jaw of a reptile that had been enlarged from just 1 cm long to at least 10x its original size.

Matt Smith pointed to a ridge running along the skull. “You can see things like the tunnel running through there…” he said. This type of detail is brought out through the printing process.

It also turns out the 3D printed jaw that Matt Smith was holding was from a Tuatara. Now found only in New Zealand, these creatures were common during the Triassic and beyond. They are like lizards, but with less flexible jaws and fused teeth that allow them to chomp down on and chew their prey. 

Finding Triassic Tuatara-like fossils in the park provides a useful link to the origins of lizards. “Lizards replace them,” explained Matt Smith, “It was like this ecological arms race.” Two reptile groups pitted against each other for survival. 

When you walk through the park today, all you see are lizards, but “they are here now because of this struggle that occurred 220 million years ago,” said Matt Smith.

And with that, Matt Smith shooed us out to enjoy the rest of our day. We were just getting started.

3D print of Tuatara jaw.

Heading Back in Time

With my brain crammed full of information, it was finally time to head out into the park. We said our goodbyes to Matt Smith the paleontologist, and Dorenda, Matt Walters, and I, hopped in our cars to begin the 28-mile drive through Petrified National Forest Park.

Heading south, we drove past the Painted Desert and pulled off for a quick stop at the Blue Mesa Member of the Park to look at some petrified wood.

“The youngest part is 209 million year ago, up where we first started,” explained Dorenda, “in the Painted Desert with the red badlands.” 

Now we were looking out at 217-million-year-old badlands of greys, blues, and greens. With puffy white clouds dancing across the otherwise expansive bright blue sky and casting shadows, the view was breathtaking. 

The view over the Blue Mesa badlands.

Keystone Arch

Hidden amongst the bentonite clay hills, were petrified logs of various sizes and shape—each also uniquely colored.

“The theme of this park is erosion, erosion, erosion,” said Dorenda, as Matt Walters led us out into the colorful environment. It is through the action of water and wind that the petrified logs that the park is famous for are revealed over time. 

Matt and Dorenda stopped in front of one of these logs that arched its way from one side of a small gully to another.

“This is a really special petrified log,” explained Dorenda, “this one is called Keystone Arch.” Aptly named—the single log was several pieces—held together by touch points between each piece. It was beautiful, but temporary structure. The process of erosion, already acting day-by-day to bring the arch down.

Keystone Arch.

Distinct Species

I asked Dorenda and Matt Walters if they knew what species of tree Keystone Arch was made from. They told me there was no way of knowing without looking at the cellular structure. Most of the tree species, with a few exceptions, are too difficult to identify without this level of detail. “There were over 1,000 species,” explained Dorenda.

Of course, some are more commonly found in certain locations. The north end of the park is called “the Black Forest,” for example, and the petrified wood there tends to be darker in color because of differences in fossilization.

“It is just like forests today,” said Matt, “different trees in different areas.”

Petrification

Scattered around the base of Keystone Arch were several pieces of petrified wood of various sizes and colors. This, of course, begs the question: “Why is there so much petrified wood in the park?”

Dorenda explained that during the Triassic Period the Petrified Forest would have been on a large supercontinent called Pangea, very tropical, and very wet—with many freshwater streams, swamps, and lakes—and of course lots of trees, some over 200 feet tall.

This combination of trees and water meant that many trees after death were toppled, as streams undercut their banks. These dead trees, often stripped of branches and bark, might then be transported downstream, collect in areas where water slows, and become buried in sediments where decay is inhibited.

“This area would have been a converging of waterways, and just a big damming of logs,” explained Dorenda.

A piece of colorful petrified wood found near keystone arch.

Colored Stone

Blues, reds, oranges, yellows, purples, and blacks—a palette of colors can be seen in each petrified log. The colors develop in the log next in the petrification process as mineral-rich groundwater travels through the logs. 

The petrified wood is mostly quartz minerals or silicon dioxide. For this reason, “you need silica for petrification,” said Dorenda. In other words, you need volcanic material.  Since there is not much of a history of volcanism in the area, much of the material was blown in from the west during the Triassic.

“Silica adheres to organic cells,” Dorenda went on, so as the silica-rich water percolates down into the earth and reaches a buried log, it enters the wood and stops. The silica alters the wood into opal, replicating its features, and eventually transforming the wood into crystalline quartz over millions of years. 

And the colors? “As the silica solution goes through the earth it picks up minerals,” Dorenda explained. Pure quartz is colorless. It is minerals like iron oxide or manganese that are responsible for the kaleidoscope of colors present in the stone.  According to Dorenda, iron oxides can create colors from yellow to red, even purple depending on the level of oxidation. Manganese creates dark woods from purple to jet black.

Some of the logs found in the Petrified Forest look a lot more like wood than stone. These logs, Dorenda explained, would have started to decay early on–creating inorganic cells that the silica dioxide won’t adhere to—resulting in weaker, lighter permineralized logs. 

Matt and Dorenda had me hold a piece of each type of log in my hands so I could feel the difference in weight. Both felt heavy like stone, but the agatized stone was a bit heavier. “One cubic foot of agatized wood weights 160 lbs,” said Matt.

At this point, Dorenda, Matt, and I navigated our way through the badlands and back to our cars to continue our journey through geological time.

Pieces of permineralized wood scattered on the ground.

The Flattops

I followed Matt and Dorenda further south into the park, before reaching a small pullout adjacent to “The Flattops.” Here we met up with fossil preparator and paleontologist, Diana Boudreau, for the main event—a hike into the badlands.

After some quick hellos and grabbing our gear, we got moving right away.

I looked out onto the unmarked terrain. Like our earlier stop, there were flat topped mesas and rolling hills—only this time in shades of grey and red brown. Despite the similar feel of an alien landscape, this section of the park marks a different time frame from our early stop at Blue Mesa—moving us forward in time to about 213 million years ago.

“The Blue Mesa region is mostly composed of the Sonsela Member. Here, we see the Flattops Beds of the Petrified Forest Member sitting just on top of Sonsela,” remarked Diana as we descended from the road into the backcountry.

“The Flattops,” as seen from the road.

Older than Dinosaurs

Making our way cross-country, with Matt in the lead, I asked Diana to tell me more about the park’s geology.

“So, the whole park is Late Triassic in age,” Diana began, and represents a range of time from 208 to 228 million years. “That is most of what is exposed here.”

Most of the fossils found in the park are not dinosaur—a common misconception–she added. Instead, they are from a much older, larger group of reptiles called archosaurs. Distinguishable by differences in ankle and hip bones, archosaurs are the Triassic ancestors of many later lineages, including birds, crocodiles, and dinosaurs.

Geological Members

With the main fossil bearing members including the Blue Mesa, Sonsela, and Petrified Forest members, Diana continued.  Each member is distinct from the others based on certain traits, like depositional environment. 

“We will be crisscrossing between Sonsela and Petrified Forest,” Diana remarked—moving between 216-million-year-old deposits of cross-bedded sandstone and 213-million-year-old mudstones and sandstones. Our footsteps dancing back and forth through time.

Cryptobiotic Soils

Matt set the pace, while Dorenda, Diana, and I followed closely behind. Watching our footsteps along the way. One of the first lessons for backcountry travel is to watch your step. Not only are there hazards to look out for, but crypotobiotic soils to protect.

A cryptobiotic soil is a dark soil crust that is formed by a suite of organisms, like fungi, lichen, bacteria, and algae, over long periods of time. These organisms are “the first biologic that grows in a sandy, arid environment,” explained Dorenda and they build up the soil in such a way that it benefits plant life and prevents soil erosion.

However, one misstep and 50 years of microbial work can be completely dismantled. Dorenda and Matt pointed out some cryptobiotic soil growing near a plant. It looked a bit like moss growing on a rock, but darker.

All the wiser, we side stepped the growing mat and continued on our way.

Cryptobiotic soil found on the hike.

Human “Footprints”

As we hiked, our footprints marked our path across the desert—a path that would later be washed away with the next rainfall. However, these were not the only signs of human presence during our walk.

Not too long after finding the cryptobiotic soil, we passed by a couple of pottery sherds —archeological artifacts of human habitation hundreds of years old—the first of many.

One of the first pottery sherds we passed during our hike.

A while later, we came across a surveying “benchmark”—a point of reference for mapping. The date on the patinated copper surface read 1921—100 years prior to our hike across the desert; placed by the U.S. General Land Office Survey before USGS existed. One-hundred years ago the park was newly established and the first phytosaur fossil found in the area was being described. 

These were early human “footprints”—impressions of a past that exists in clues and signs.

A survey marker, or benchmark from 1921.

Geology

For Diana, Matt, Dorenda, and I time passed quickly—both literally and figuratively—as we walked over the undulating hills.

At one point, Diana stopped abruptly. “This is a nice vantage point to talk about the geology here,” she remarked.

Looking out to our left was a steep, eroding cliffside with horizontal bands of varying shades of brown. Diana directed our attention toward these bands. “This view shows a lot of the different flattops beds,” she said pointing.

Diana then went on to describe each layer as a numbered unit starting with a section of sandstone at the top, followed by alternating layers of mudstone and sandstone. She explained how each deposit would have been laid down by a braided stream system—with sandy material deposited in the stream bed and more silty/muddy material along the banks.

The units we were looking at represented a time spanning about 1 million years from 213 to 212 million-years-ago. Starting with Petrified Forest Member Flattops beds at the top and the more rounded Sonsela Member at the base of the cliff. Each layer was thick, indicating a water rich environment over the 1 million years, but would have differed in the type of watery environment and the organisms that lived in that place at the time.

Diana pointing out the geological units.

Petrified Peat

As we continued hiking atop the Sonsela hills, Diana and I chatted, while Matt and Diana led the way, eventually stopping near what looked to me like a short fence made of solid rock—a line of stone stuck out of the ground vertically. 

“This area here is what we called silcrete,” said Dorenda. “Remember when we talked about the petrified wood and we talked about those giant logs? And how as the water took its toll the branches and bark and everything would be gone?” she asked me.

Well, according to Dorenda, silcrete is the result. Waterways collect all the partially decomposed wood bits in one place where they undergo the same process as the logs and are petrified.  “It is kind of like petrified peat,” Dorenda stated.

Standing on End  

Of course, usually, silcrete is laid down in horizontal layers. The silcrete here was vertical. Making the spot a bit of a geological mystery, as there is no sign of faulting that might normally turn rock on end. 

“It is theorized that there was some sort of pressure that pushed it up through the sandstone, filling in the gaps,” Dorenda told us.

“You will find silcrete through the whole park,” said Matt, but vertical layers like this “can only be found in two places.”

Petrified peat, or silcrete, creating a vertical “fence.”

Wandering the Wilderness

After ample time spent taking pictures of the silcrete anomaly, the four of us continued our hike under blue, cloud-spotted skies.

With Matt Leading the way, Dorenda, Diana and I hung back and discussed a variety of topics from career choices to canyoneering, before the conversation shifted to the preservation of natural and ecological resources.

The Petrified Forest National Park was one of the first National Parks to have land set aside as designated wilderness in 1970. Wilderness is the highest form of protection public lands receive—restricting access to those on foot and limiting human impact. 

However, the Petrified Forest is unique in that it offers hikers and backpackers the opportunity to explore outside of a designated trail. In fact, it is encouraged, as there are no major trail systems in the park.

Of course, “leave no trace” principles still apply and often require extra consideration, especially in a place like Petrified Forest where archaeological artifacts and fossils are abundant. Even taking pictures requires special consideration to preserve the location of unique places that might draw crowds that may end up impacting the park negatively. Dorenda and Diana both expressed concerns about protecting artifacts and other special place locations. “It’s a weird line,” said Dorenda, but an important one, not only for the resources being protected but often the safety of visitors to public lands as well.

Matt leading the way, leaving nothing but footprints.

An Eye

By this time, we were at least a couple of miles into the wilderness and the road where we started was a distant memory. Lost in the vastness of the wilderness (but not really lost thanks to our guide Matt), I asked Dorenda and Matt if they had any advice for those visiting the park, or any other natural place, for the first time.

Dorenda spoke first. “Take the time to see the micro and macro view,” she said. “Do a 360,” she suggested. She told me that she tends to keep her eyes to the ground. It takes deliberate effort to stop and look around.  But taking in both views will help you better appreciate all aspects of the park.

In addition, “I think you develop an eye for things,” she went on. Whatever you are looking for, whether it’s fossils, petroglyphs, or something else, if you learn what to look for you get better at finding it. Look for contrasts, different colors, textures, and size and that will help you

A Guide

Matt had a different take. “It is all being passed down,” responded Matt. Learning about a place from others that know a lot more than you do can really help enrich your experience. 

“We had two mentors. They taught us a ton because they had to teach us the hikes,” he continued.  “We were like kids in a candy store because we were learning so much.”

Taking it further, Matt recommends sharing what you learn. “The key this for us is to pass it onto people,” he remarked.

Having spent more than half the day with Dorenda and Matt, I was able to see this key in action. And let me tell you, they are well practiced.

Candy store, Matt? More like Wonka’s Chocolate Factory!

Matt and Dorenda Walters leading the way.

Finding Fossils

One thing to know about Matt, is that when he slows down on a hike, it is time to look around.

It was getting near lunch time, and we had picked up the pace in an effort to make it to a lunch spot Matt and Dorenda suggested, so when Matt stopped abruptly, we knew there must be something interesting nearby.

Diana spotted it right away—a fossil! Laying on the dry desert floor was a small fossil of a bone, about the size and length of a snickers bar. It had a crackled texture and was broken in one place.

“I think it is a phytosaur rib,” said Matt.

Diana looked closely and agreed that “it was the right size for a phytosaur.” Definitely a long bone—either “a process from the vertebra or could be a rib,” she said.

She picked it up and we looked closely at the fossil, pointing out the cellular structure visible in fossil bones before laying it back down.

I tried to imagine a large reptile sitting in a swamp waiting for its prey, but it only made me think of my own lunch waiting ahead of me.

A long bone fossil found during the hike.

Keep Looking

Continuing along we saw several more artifacts laying on the desert floor—including a piece of a corrugated pot and another fragment of a vessel with a small hole in it.

Before long we had reached our lunch site, but we weren’t “allowed” to eat just yet. Matt said that I would need to “earn my keep first,” as there was another artifact nearby and it was my job to spot it.

After several painful minutes of trying to spot what I thought would be pottery or a fossil, Matt guided my eyes to a faint figure inscribed onto a dark colored rock—a rock I had been staring at for a good three minutes, at least.

Petroglyphs

The petroglyph in front of me was the impression of an animal of some sort—carved into the dark desert varnish growing on a rock.  The image was faded—the result of time passed—as the bacterial growth responsible for the varnish was starting to repopulate the etched-out areas.

“These are probably over 1000 years old,” said Matt regarding the petroglyph.

Looking closely at the rock you could see small dimples formed the petroglyph impression. Matt explained that the petroglyphs would have been chiseled into the rock, probably using a tool made from petrified wood and a hammer stone.

Having “earned by keep,” we found some other stones to sit on and enjoyed a leisurely lunch basking in the warm desert sun.

Can you see the petroglyph?

Artifact Delights

After lunch, things really got moving, as we drew closer to a larger archaeological site in the area.

Diana spotted a small unionid bivalve shell, or mussel, from one of the many species common in the area.

Matt Walters also led us past a vertebra fossil and a collection of other fossilized bone fragments, as well as several fragments of broken pottery, before reaching the piece de resistance—the site of several ancient Puebloan pit houses. 

The mussel shell Diana found during the hike.

As Matt, Dorenda, Diana, and I neared what I later learned was a pit house village, we started seeing more pottery fragments, as well as several other archeological artifacts.

The pottery was of various colors and textures and used a variety of design elements—there were white and black pieces, fragments of grayware and corrugated pieces, as well as some decorative edges and unique shapes. Dorenda explained how some of the pottery would have been traded into the region, while other pieces were likely made by the local people.

Flakes from arrowhead and other tool-making also scattered the ground in colorful abundance. It was fun to pick out some favorite pieces to admire before moving to the next.

One of many pottery shards found during the hike.
`Flake of petrified wood found during our hike.

Pit House Village

The density of the pieces continued to increase as we neared a few mounds of rocky earth—we had arrived at the pit houses. Matt Walters estimated that there were probably three dwellings in the area. And what a view! I guess the old adage “location, location, location…” is more ancient than I thought.

The pit houses themselves would have been built by stacking rock vertically and digging down into the earth. Then a roof would be fashioned out of whatever materials were available. There would have been a garden of squash, beans, and corn in the area and probably some storage pits as well. Though the pit houses were permanent dwellings, they were often only used seasonally.

Looking down on the pit house site. What a view!

Pit House Treasures

The areas around the pit house ruins contained many more archaeological treasures. Dorenda showed me a rounded stone, about the size of a human hand. “It’s a hammerstone,” she said. “You can tell it has been used because it has chips in it.” The stone was heavy in my hands.

We also saw several large grindstones comprised of a large flat stone, called a metate, and a smooth stone with a shape suitable for grinding.  The metate was ground down and smoother with use.

We also saw several unique pottery pieces, some so fragile that we avoided picking them up, including a small piece that had a handle and looked a bit like a ladle.

A hammerstone
A grindstone.
One of several metate found during our hike.
Fragment of pottery with handle.

Weather or Not

After visiting the pit houses, we slowly curved our way back toward the cars. There were still plenty of artifacts to see, including several intact arrowheads and many more flakes of petrified wood.  Matt Walters led us around to all the fascinating finds as we hiked.

We passed by another petroglyph site, before heading into a canyon between two flat topped Mesas. Though dry as a bone at the time, Matt and Dorenda told Diana and I how once they had found the canyon impassable from flood waters. “This was a roaring river,” said Dorenda. Though rain might not be frequent in the area, flash floods do occur.

The lasting influence of rain could also be seen on the canyon walls—gullys and rills marked the paths of past water events. There were also large holes at the base of some of the hillsides, created by the movement of water along paths inside the Earth that widened over time.

Looking up, the hills wore sandstone caps—created by the weathering of the softer mudstone below.  Giving the place an overall hoodoo-like quality.

Sandstone caps on top of softer mudstone.

The CCC

We continued to follow the dry riverbed into the canyon. Large jumbles of rocks lined our path most of the way.As I considered these large fallen stones, Matt Walters slowed his pace again. Sitting amongst the rocks was a long piece of wood.

Matt Walters inquired— “Who had a big impact on the park?” He asked.

“The CCC,” he said, after some deliberation.

According to Matt Walters, the Civilian Conservation Corps were in the park from July 1934 to 1938. The long piece of wood was an artifact of that time. “The CCC had a flagstone quarry,” Matt explained, “we think this is part of the quarry.”

In addition to the quarry, the CCC built the Painted Desert Inn and dug a 16-mile irrigation system in one years’ time.  With current regulations and protection for archeological and paleontological resources stricter, it took three years to replace that same waterline in 2016.  Matt Walters chuckled at the irony.

A long piece of wood.

My Own Eyes

Just a bit further down the wash, I had my own fun. Hidden amongst the rocks, I made my first solo fossil discovery—another freshwater bivalve shell lay on the ground.  I called out to the rest of the group to share in my triumph. Then I took a few pictures of the fossil shell before placing it back on the ground for another to find. 

 A Few More Petroglyphs

As the sun sunk a bit lower in the sky, we entered the last leg of our journey which brought us to a couple more incredible petroglyph sites.

At one site there were several large stones decorated with at least a dozen figures, ranging from bear claws of various shapes and sizes to what looks like a coyote. Human figures were also displayed on the slab that was probably about as long as I am tall. 

Petroglyphs covered several large slabs of rock at this site.

The last petroglyph site was a bit more of mystery. Here a large rock was marked with several dot-arrays, a couple of straight-lined figures, and a set of zig zags as a border. The whole display seemed to be conveying some sort of information, but what? 

Matt, Dorenda, Diana, and I all puzzled over it, offering hunches and second guesses as to its meaning, before moving on.

Petroglyph with mysterious message.

Stone Tree

Just before hitting the road and leaving behind our backcountry adventure, I noticed a lone piece of petrified wood sitting quietly on the brown, cracked Earth.

Perhaps I was developing “an eye” for this unique desert environment because I felt drawn to it. So much so that I snapped a quick picture.

A few moments later, after goodbyes and well wishes, I was back in my car, driving the lonely road to my home for the night.

Last picture of the day.

A Snapshot

The picture of the log is the last one that I took that day. Looking at it now, I still feel its call. A call to a time before the dinosaurs—to swamps and rivers hidden in a now desert landscape. To a time where people lived in pit houses and hand-crafted stone tools and beautiful pottery. A call to modern-day adventures and new friends. And finally, a call to return to this place someday in the future—to remember, while discovering the past, all over again.

Dorenda and Matt Walters are long-time volunteers for Petrified Forest National Park, guiding park guests on fabulous cross-country hikes each weekend. Diana Boudreau is a paleontologist and fossil preparator at Petrified Forest National Park. Matt Smith is the long-time museum curator for Petrified Forest National Park.