Hike with a Meteorologist

View from upper Rodney Falls looking down.

The sky was crying out large splattering raindrops quicker than my windshield wipers could sweep them away. Traffic was usual for a Sunday morning in Portland, Oregon, as I headed north on I-205, and eventually into Washington state. My destination: Hamilton Mountain Trailhead. My purpose: to meet up with meteorologist and fair weather hiker, Steve Pierce, for an afternoon hike and interview. 

But the rain was hammering and I was beginning to wonder if this thing would happen. Steve had told me straight-out-of-the-gate that he didn’t hike in the rain, so he probably wouldn’t be up for a downpour. 

What is up with the Weather?

Weather: atmospheric conditions present at any given time, on any given point on Earth. Such a simple definition for such a complex, life-altering phenomenon. The weather can turn a good day into a bad one, build up our snowpack, exacerbate forest fires, leave us exposed, fearful even. And in my case, possibly defer a hike?

The weather was doing its thing. So I drove and crossed my fingers (figuratively), tracing the path of the Columbia River to my right as I headed east on Highway 14. By the time, I reached the trailhead, conditions weren’t looking much better. By the time Steve pulled up in his little red convertible, we were still in the thick of it.

Windows rolled down and raindrops falling all around, I met Steve Pierce for the first time, sitting side by side in our vehicles. After some quick introductions, he pulled out his phone to check the weather radar. “Another 15 minutes or so and the rain would pass,” Steve told me with confidence. 

A Career

The hike was on. We just needed to wait out the rain for a bit. So Steve slid into my car and told me how he got started in the world of weather.

Beginnings

Inspired by events like the 1980 Mount Saint Helens eruptions, from a young age Steve was interested in weather. “I was fascinated with which way the ash clouds were going,” said Steve.  When ash fell in the Vancouver-Portland, Steve was only seven but he remembers the experience vividly.  In particular, he recalled spending time with his brother sliding around on their bikes on a slurry of ash and water that felt akin to snowfall. Coupled with a couple of major snowfall events earlier that same year, the appeal of meteorology hung heavy in Steve’s young mind, like (dare I say) rain in a cloud-ready to fall. 

Self Taught

Throughout his youth, Steve studied the weather on his own. He read every book he could get his hands on, including every meteorology book at his local library. Later, when the internet was a thing his studies continued. 

Early on he also tried his hand at forecasting. In 5th grade, he started writing a weekly weather forecast for his elementary school’s parent bulletin. At home, Steve would broadcast the weather from the family fireplace hearth through his Mr. Microphone. He also had a weather station and was a devoted weather spotter for Channel 8. By 9th grade, he found himself on the cover of the regional paper (The Columbian) in an article titled: “Teens Career Forecast is Clear.”

You’re Hired

Fast forward to 2014, after attending Mt. Hood Community College (1994) and Washington State University, Vancouver (2001), a family, a prior communications career, and“fever” for investing in real estate, Steve finds himself at a crossroad in his life. At this point, Steve is an active member and President of the Oregon Chapter of the American Meteorological Society (AMS), but otherwise had made his career outside of meteorology, when he gets “the call.” It was local news anchor, Jeff Gianola, who remembered him from 20 years prior, when Steve was a college intern, with a job opportunity: Meteorologist for KOIN 6 News. Steve is instructed to: “get on a suit and come to the station” if he is interested in this rare opportunity. Adorned in basically the only formal suit Steve owned at the time, he goes into the station for a mock-up weather forecast. He is hired on the spot!

Time to Hit the Trail

This brings us to today.  Steve and I had been talking for several minutes in the car at this point so that the windows were fogged up and the air a bit warm and humid. The rain, on the other hand, had finally stopped. I suggested that we hit the trail. 

Steve Pierce at the Trailhead

The Hike

  • Trailhead: Hamilton Mountain Trailhead
  • Distance: approximately 2.5 miles round trip to Rodney Falls
  • Elevation Gain: approx. 450 feet elevation gain
  • Details: The trail to Rodney Falls is an easy out and back hike. Those that want more of an adventure may continue to the top of Hamilton Mountain for a total of 7+ miles and 2100+ feet elevation gain . There is ample parking at the trailhead and restroom facilities available. Parking requires a WA Discover Pass.

Basic Science 

Steve and I started our hike with a couple of photos before making our way up the forested trail. The sun was coming out and the only water falling around us was from the surrounding vegetation.

“See I promised it was going to be good,” Steve smiled. 

As we moved at a quick pace, I asked Steve to tell me about what it takes to understand the weather. What does it take to forecast the weather? He said two things: 1) basic science and 2) interpreting computer models. 

“Anyone can be a weather enthusiast,” replied Steve, “but sort of where-the-rubber-meets-the-road is really understanding meteorology on a scientific level. How are weather patterns created? How do they evolve?”

He went on with an example. “We live in a part of the country where the Columbia River Gorge is huge in determining our weather. Without the Gorge, we would rarely get snow in Portland.” He explained that the Cascade mountains create a barrier that separates the cold, dry continental air mass that sits over the continent during the winter. The Gorge is a gap in that barrier. “The gap is how we tap into that cold air.”

And this is where we get into some of the basic science of meteorology. Cold air is dense and tends to sink, and warm air rises and expands. Air also moves from low to high pressure.

So, if we look at the science, it makes sense that in the winter, the cold air from the east would be more than happy to rush through the Columbia River Gorge following the pressure gradient, on its way to the relatively warmer ocean. Then, if there is some moisture on top of that, bring on the snow! 

Computer Models

In addition to knowing the basic science, “interpreting computer weather models” is essential to meteorology, especially in modern times. 

Computer weather models take observations from weather stations and weather satellites and assimilate them into a 3-D grid. They then use mathematical equations that describe the physics of atmospheric variables, like pressure, temperature, wind, and moisture, to make predictions about future atmospheric conditions. Thus, weather models depend on both the equations used and the observation. The better both of these play together the more accurate the forecast. 

Steve shared that the models he prefers are a set found at the University of Washington, developed by “long-standing atmospheric science professor, Dr. Cliff Mass.” NOAA also has some high-resolution models too. 

Computer models have truly revolutionized weather forecasting, making accurate forecasting possible. This is the “leading-edge technology” we are talking about. Available to all, these models make forecasting the weather a breeze (pun intended) compared to the past. The Labor Day winds and forest fires, for example, were forecasted to the hour by weather models.  

Again, most of these models are available for those who are interested. 

Why Weather?

And who isn’t interested? 

The weather has always been kind of a big deal—affecting our past, present, and future. 

As Steve puts it, “Weather affects everybody. It affects your everyday life, your job if you work outside—agriculture, for instance.” The weather today can also affect the future. 

For example, the amount of winter snow in the Pacific Northwest, predicts future water supply in the region. All that snow that dumped down in the valley and mountains a few weeks ago wasn’t just weather for now, but water for later. We even saw remnants of the snowfall during our hike along the trail. 

“I just saw the latest snowpack for Mt. Hood,” said Steve, “and we are at 100-110% of normal.” And that is good news for the upcoming summer water needs across the Pacific Northwest.

Snow on the trail!

A Tale of Two Clouds

The forested trail continued into a short clearing where I could feel the sun warming my skin. In his cavalier way, Steve noted this change in weather. He had promised it would be nice.

To the right, there was a bit of a view out toward the river and a basalt cliffside, but my attention was drawn to the clouds overhead. I asked Steve, “What kind of clouds are these?”

“These are cumulous,” replied Steve. 

Clouds are a personal favorite of mine. And growing up in Oregon, there never seems to be a shortage of clouds to look up at. So of course, I needed to know more, so I asked Steve to elaborate. 

As he explained it, there are many types of clouds, but two of them are the most interesting to him: cumulus and stratus.

Cumulus

Cumulus clouds, like much of what we saw during our hike, are convective. What does that mean? It means that the droplets of moisture that make up the cloud are heated up by the surface of the Earth and rise up. Just like a pot of boiling water, cumulus clouds “puff-up” as they rise, like bubbles breaking the surface. Cumulus clouds are also a sign of an unstable atmosphere and that severe weather may be on its way.

Stratus

Stratus clouds on the other hand are low to the ground. Unlike cumulus clouds, stratus clouds are stable and flood in from coastal areas in thick layers of moisture.

According to Steve, “Stratus are unique to the west coast.” During the summers, the air above the ocean is colder and denser than inland. This colder air can be held at bay during the day, but by 6 p.m., as Steve puts it, “mother nature’s air conditioning kicks in,” and the cooler coastal air drifts inland through gaps in the coast range, including the Columbia River Gorge. Oftentimes, this cool onshore flow will also bring in stratus clouds, filling-in the valley overnight. 

Viewpoint with some “dissipating” cumulus clouds

Clashing Air

While cumulus clouds continued to drift east through the Gorge, Steve and I continued our hike in the same direction, leaving the cloudy peek-a-boo view behind.  

Moving swiftly up the trail now, my mind was racing, as I attempted to assimilate all that was being said. One word stuck in my mind: “unstable.”  Steve had thrown this word around a bit. If today was “a bit unstable,” as Steve suggested, what does that mean? So, of course, I asked. 

“An unstable air mass, simply put, is very buoyant air,” Steve replied. “Unstable air is usually what follows a cold front. And a front is the clash of two air masses, warm air ahead and cold air behind.” The clash between the cold and warm air builds up clouds as the colder, denser air mass replaces the warmer, less dense air mass, resulting in cooling and condensing of water droplets that form clouds and can produce rain.

A Big Difference

Steve further explained that the bigger the difference in temperature between the air masses, the more unstable the air mass can become. And the greater the instability, the bigger the potential for convective storms.

In the Pacific Northwest, the biggest storms tend to occur in the spring as there is more heating of the Earth by the Sun, creating more frequent and larger temperature differences between air masses and in between the different layers of the lower atmosphere, say below 25,000ft.

Extreme Weather

However, despite the Pacific Northwest’s reputation for weather in the form of rain, we do not have the extreme weather that you might see, in say, the midwest. 

In the midwest, supercell thunderstorms develop and spin up tornadoes because of the clash between cold continental air from Canada and warm air from the Gulf of Mexico. In the Pacific Northwest, this sort of instability just doesn’t happen often. The differences in temperatures over the Pacific Ocean land are just not that significant. As Steve put it: “We just have relatively cold” air.

Cold to the Core

Instead, the Pacific Northwest’s version of a twister is a much calmer cousin—the cold-core funnel cloud. A cold-core funnel (while also sounding like a new exercise routine) is the Pacific Northwest version of a tornado. Simply put, it is a vertical “rotating” column of air in the atmosphere that is especially cold, unstable, and almost always trails just behind the recent passage of a cold front.

“After a front has passed, we typically have westerly flow at the upper level of the atmosphere, and a southerly or southwesterly flow up the Willamette Valley”, Steve explained. “Just enough wind sheer between the two, along with geographic enhancements in the valley, and you can spin up a horizontal column of unstable air and then turn that column vertical due to the wind sheer present.” Hence, a tornado.

If these cold-core funnel clouds touch down, though rare, they can cause EF-0 or EF-1 tornado damage. 

In fact, according to Steve, the deadliest tornado west of the Rockies hit Vancouver, WA on April 5, 1972. Rated as an EF-3 (F-3 back then) tornado, winds were 165-210 miles per hour. Six people died that day in Vancouver when a supermarket and bowling alley roof collapsed. The roof and walls of near-by Peter S. Ogden Elementary School were also damaged.

Large Bodies of Water 

However, deadly storms are anomalies in the Pacific Northwest. “We don’t really have that kind of volatile weather,” explained Steve, “because we have the Ocean.” The Pacific Ocean, as is the case with any large body of water, is slow to heat up and slow to cool down, making it an excellent moderator of climate. 

In the summer, when the inland starts to heat up, the Ocean lags behind. In the winter, temperatures drop inland, but the Ocean retains its heat. The result is cooler summers on the coast and warmer winters.

The moderating effect keeps Portland and other Willamette Valley temperatures from reaching extremes as well. Hiking in the Gorge, even in early March, is pleasant and cool, but not too cold (at least relatively), all thanks to the Ocean to the west. 

Weather Patterns

One of the first things Steve talked about when I asked him about his interest in meteorology was the importance of weather patterns. 

Here is what he said:

“What I am most passionate about with meteorology is I know things happen in cycles. In other words, when we saw the big snow event coming just a few weeks ago, I thought back to previous events where I knew this happened in the past… I remembered I have seen this pattern before. Weather is a set of different patterns that come and go.”

In Oregon, typical patterns of weather result in a climate that is warm and dry in the summer and cool and wet in the winter, and overall pretty moderate. But, on the other hand, we can also predict deviations from the norm, like the winter storm Steve explained earlier. 

Much too Hot

Another example of these patterns of predictable deviations occurs in the summer. 

Typically during the summer in Oregon, there is a ridge of high pressure over the ocean that blocks storms between June and October, while prevailing winds from the northwest, carry cooler ocean air onshore as a high-pressure system spins clockwise in the northern hemisphere.  This is the norm.   

However, there are also usually a few days of 90+ degree days, where things don’t cool down.   In Oregon, you can expect a couple of brushes with high-temperature days in the summer—it’s a pattern. This occurs when a low-pressure system from the Southwestern United States creates a thermal heat trough that moves into western Oregon. This turns the airflow counterclockwise, directing the winds offshore toward the coast. Not only that, but the downslope flow of air of the Cascade mountains results in “compressional heating,” a thermodynamic process that heats up the Willamette Valley of Oregon even more. 

A Numbers Guy 

Now one thing that became pretty obvious early on is that Steve is a numbers guy. He is self-professed “good with numbers” and bad with names. I am not great at either so I was impressed how easily he rattled off dates when I can barely remember my own kids’ birthdays.  

“Hottest temperature in Portland: 107o F on August 10th, 1981,” Steve rattled off as we marched along.  “Coldest temperature in Portland: -3o F on February 2nd  1950.”

After tip-toeing our way through a mud patch, he went on:

“Average temperature in Portland is about 53o F… about the same temperature as the Pacific Ocean… It balances out.”

Superlatives

Steve continued with the superlatives:

  • Wettest months: November & December (a near “tie” at 5.60” and 5.40” respectively).
  • Most active weather (thunderstorms): March through May
  • Foggiest month: October
  • Biggest change in temperatures: Also October. “With an average range of high’s from 71o F on the 1st to 58 o F on the last day of the month.”
  • Best Month: October (in Steve’s opinion, due to the change of season and falling of leaves)

Chasing Waterfalls

As Steve and I waxed poetic about the beauty of fall foliage in the Pacific Northwest, the rushing sound of water filled our senses. We were nearing the base of the middle of Rodney Falls located on Hardey Creek. Here, we decided to stop the interview for a bit to check out this fun natural feature.  

Steve and I first took an upper trail to the left to get a close view of one of the tiers of the multi-tiered waterfall, before heading down to the base of the fall where you cross over Hardey Creek to continue the hike. With so much snow and rainfall recently, nature’s tap was on full blast and the falls were spectacular. 

Rodney Falls

A Cliffside View

Just beyond the falls, Steve and I continued our hike a few hundred feet up to a viewpoint looking out at the Columbia River Gorge. From here you could see the north-facing wall of the canyon rising up on the Oregon side of the Gorge, still covered in snow. 

Looking out at the basalt rock cliffs, Steve and I shifted our discussion toward the landscape. (Literally, as we faced the canyon wall.)

“Geographic features have an effect on weather patterns,” Steve said. The Gorge plays a part in shaping weather patterns in the valley. The Pacific Ocean moderates temperatures all along the western margin of the country and beyond. 

View of the Columbia River Gorge just above Rodney Falls on the trail

Rain Shadow

There are many geographic features that influence the weather in any location, but one that is particularly important to the Pacific Northwest is our mountain ranges. The Coast Range and the Cascade Range both run north-to-south in the region, creating barriers for winds and weather. 

Mountain ranges, like those in Oregon, create what is called a rain shadow—a dry landmass on the opposite side of the mountains from which the air is flowing. As cool, moisture-rich air from the coast travels up each mountain range’s western flanks (also known as the windward side), the air cools and moisture condenses forming clouds and precipitation. Once the air makes it to the eastside (also known as the leeward side), the air has been wrung out like a sponge. As the air descends further it also heats up and evaporation increases. 

In Oregon, this means that it gets drier as you head east across the state. According to Steve, the coast averages 60 inches of precipitation, Portland 37, and Bend less than 10. 

This also means that in the east you experience a very different ecology than in the west. Instead of the shaded, wet Douglas-fir forest, we were hiking through, hiking the east would take you through open, dry shrub-steppe habitat, or perhaps a Ponderosa Pine forest. Hiking in the east means applying a lot more sunscreen and packing more water. 

Douglas-fir trees line the muddy trail.

Sunrise & Sunset

After sufficient time at the viewpoint, Steve and I decided it was time to turn around and head back. And as the sun was setting on our time together, I couldn’t help but ask Steve about literal sunsets—can we use the weather to predict the best sunrises and sunsets?

Yes you can! 

“The first thing I would do is look at satellite imagery,” said Steve, “…nothing coming off the coast is a good sign.” If there is a ridge of high pressure over us that shuts down the onshore flow that equals a good sunrise or sunset. 

Steve also recommended winter, as the best time to see sunsets. “The atmosphere is more mixed up,” he explained, “so you get really clear air.”

He also warned that sunrises can be more tricky though because of the morning stratus clouds that move off the coast and through the valley at night. These clouds can put a damper on a good sunrise. Of course, when in doubt you can always head east. The clouds can’t make it inland that far. 

The Road Home

After Steve and I finished our hike, I climbed back into my car and began the long drive home west through “the Gorge” and south into “the Valley.” 

On my way home the pleasant weather turned from partly cloudy to stormy in what felt like an instant. Hail fell from the cumulus clouds overhead and I am pretty sure I saw a flash of lightning off in the distance. Steve was also stuck in Washougal, WA. by a torrential downpour, thunder, and lightning.

The stormy weather was short-lived, however; and as I pulled into my neighborhood the sun was setting, turning the clouds a soft shade of pink. Seeing the painted sky brought me a sense of calm following the chaos of the weather-filled day. I thought, “How very appropriate; this is what it is like to hike with a meteorologist.” 

Steve Pierce is a part-time meteorologist at KOIN 6 News based in Portland, Oregon. He grew up in Vancouver, WA where he first developed a passion for weather. Steve studied Television Production at Mount Hood Community College and Business and Communications from Washington State University, Vancouver. He is also the President of the Oregon Chapter of the American Meteorological Society since 2012.

Hike with a Lichenologist

One of many cyanolichen found near the creek during the hike.

Equally unassuming and complex—lichen often go undeservedly unnoticed. Open your eyes to these delightful organisms, and an entire world unfolds before you. And here’s a shock, this world has been there all along hiding in plain sight. On the branch of a tree, or on the ground, or attached to a rock—lichen are everywhere. 

Heck, spend enough time in the “world of lichen,” and they may even change your life. At least that is what happened to lichenologist Joseph (Joe) R D Meglio, whom I met on a cold, wet morning in January to hike and talk lichen. 

The Hike

  • Trailhead: Baker Creek Trailhead, McDonald Forest (Corvallis, OR)
  • Distance: varies
  • Elevation Gain: varies
  • Details: This is one of many trailhead in the McDonald Forest, which offers a variety of hiking options on trails and logging roads. Limited parking is available.

Joe’s Story

Our hike began as we walked across a bridge and onto a wooded logging road, with cold pelts of mixed snow and rain falling on our heads and shoulders. Joe and I slogged up the gravel path making small talk along the way.

After a few minute’s time, I asked: “Why lichen? How did you end up here— with a career in lichen?”

Here is Joe’s story:

The Early Years

Despite being fascinated with living things from a young age, and an early interest in fungi and lichen, Joe didn’t become a lichenologist straight out of high school.  He went to college for a term, and as he put it, “I flunked out.”

College “wasn’t an expectation in my family,” explained Joe. So it wasn’t a big deal when Joe dropped his academic pursuits, opting instead to learn a trade. Joe worked as a mechanic for a few years, and later as a high rigger for a logging company. 

Epiphany

Then one day while on a break from rigging up cables, Joe found himself amongst the branches of a Douglas-fir tree in the Mckenzie River Valley. Looking around, he noticed two contrasting worlds—a clearcut landscape in the distance, a product of the logging industry of which he was a part, juxtaposed against the miraculous biodiversity of lichen dripping off the trees in the surrounding forest. He knew at that moment he needed a change.

Joe immediately climbed down from his perch and told his boss—“I don’t want to do this anymore. I am leaving.”

And he did. Right then and there. 

A Whole New World

Not long after, Joe enrolled at the local community college and eventually made his way to Oregon State University to work with lichenologist extraordinaire, Bruce McCune. 

Joe still works in the McCune lab on lichen taxonomy, and contracts with the U.S. Forest Service on Lichen related projects. He even married a lichenologist, whom we met up with later in the hike. So, yeah, Joe is really into lichen.  

Joe with a tree branch coated in lichen

Pairing up

As we followed the road uphill, we continued to chat about Joe’s work and lichen-filled life. Then I asked him the all-important question, “What exactly is a lichen?”

Joe explained, “Lichen are composite organisms” containing a mycobiont (fungal component) and photobiont (photosynthetic partner) or two. The most common photobiont is green algae, followed by cyanobacteria. A tripartite lichen will contain both, like Lobaria pulmonaria (sometimes referred to as lung lichen), commonly found in lowland to mid-elevation forests of the Pacific Northwest. 

Whatever the pairing, a “symbiotic” relationship, or cooperation, exists between the bionts. The photobiont provides food for the fungus due to its ability to photosynthesize. While the mycobiont provides shelter for the photosynthetic partner. 

Pairing up also means physiological, chemical, and reproductive changes from the original forms of the individual bionts. Nearly nothing of the individual remains. 

Hello Lichen

At this point, Joe stopped at the side of the road, grabbed a tree branch, and pulled it down to eye-level. It was time to take a look. 

Joe began to point out all the different lichen species. “Stuff like this usnea is a chlorolichen,” he said pointing out a light green, stringy-looking lichen, “and this leafy species, Platismatia glauca is also a chlorolichen,” he continued. 

On just a single branch, Joe pointed out at least a handful of different species— each with a unique color, shape, and form.  

Several species of lichen growing on one tree branch.

Growth Forms

Which begs the question— how does one even begin to keep all the different lichen straight? I asked Joe to provide some beginner tips. 

One way to start to narrow things down, Joe explained, is by becoming familiar with the different growth forms lichen exhibit. There are three basic growth forms for lichen: fruticose (shrubby), foliose (leafy), and crustose (crust). Most species will exhibit only one growth form, and even within a genus, species typically share forms. Though there are exceptions.

When it comes to fruticose and foliose lichen (aka macrolichen), the bodies (or thalli) are organized internally into layers. Joe demonstrated this stratification by plucking up a Platismatia stenophylla species (a foliose lichen) off the ground and showing me how its ventral cortex (top side) and dorsal cortex (bottom side) differed. Sandwiched in-between, the photobionts are housed, just below the upper cortex amongst the loosely packed hyphae that make up the medulla. 

In tripartite lichen, additional structures are needed, called cephalodia.  Often wart-like in appearance, cephalodia create an anaerobic environment that cyanobacteria require “to fix atmospheric nitrogen, an important part of the forest ecosystem nitrogen cycle.”

Platismatia stenophylla

Identifying Features

Of course, growth form will only get you so far when it comes to identifying lichen. Other characteristics that are helpful include the size, shape, and color of the thallus (the main body of the lichen) and lobes (branches), as well as the shape, position, and color of reproductive structures.  There are also many specialized features, like cilia and pores (and the cephalodia mentioned earlier), that can help one distinguish between species and genera.

One of my personal favorites is that of the Usnea genus. Usnea has an inner cortex that when you pull on it, stretches like an elastic band. But beware, the band easily breaks. 

Usnea longissima

This One is Not Like the Others

However, according to Joe, even members of the same species can exhibit a great deal of variability depending on the environment in which they reside. 

He picked up a Platismatia from the trail and pointed to its frilly edges. Joe explained lichen will often exhibit “extreme dimorphism,” with fringed edges, dieback, and red spots, for example. 

Even normal seasonal changes and variability in light exposure will alter the appearance of lichen thalli. Exposure to light often darkens the color, while shaded individuals may appear pale. 

Plastismatia species with “frilly edges”

Cryptic Organisms

However, even lichen that superficially appear to be the same species, may vary substantially in other ways. As Joe described it—within the same fungal genus you might find individuals with different secondary chemistry, or an entirely different genome than you might expect.  Lichen “are very variable, very adaptable,” Joe stated.  “The closer you get the more you discover.” 

In fact, looking into the less obvious differences between lichen is a big part of Joe’s work. By looking really close and comparing the genomes of different lichens that appear similar, he can parse out different species and determine who is related to whom.  

A Closer Look

Currently, he is working on reevaluating the Sticta genus, a group of lichen that are distinguished by having cyphellae, “little windows,” usually found on the bottom of the lichen. Through his work, Joe has found that what was thought to be one species, Sticta fuliginosa s. l., is really three distinct species with different traits. “One has these really interesting lobules that are digitate,” for example.

We took a closer look at one of the new Sticta species Joe is describing—a small brown, unassuming lichen—with the proposed name: Sticta gretae sp. nov. Using a hand lens, Joe showed me the little white dots on its lower cortex, its cyphellae.

“a brown, unassuming lichen” soon to be named Sticta gretae sp. nov.

Biodiversity of Lichen 

Walking along the trail it was not difficult to find lichen. Hanging from a branch or trunk of a tree, attached to a rock, or growing on the forest floor—no matter where we looked, there was plenty of lichen to look at.  

Lichen are adapted to nearly every habitat on Earth, providing a symphony of biodiversity and plenty of eye candy. With our temperate rainforest and a diversity of habitats, it is no wonder that the Pacific Northwest is “the center for diversity for fungi in North America.” According to Joe, “there are approximately 580 genera and over 1400 species” of lichen in the region—a biodiversity hotspot.

Every Niche

However, every lichen species present in the Pacific Northwest isn’t going to be found in every location. Each lichen species is specialized for a particular niche, a particular home, and a way of life in the environment.  

Joe pointed out a Peltigera species growing on the forest floor. “These are terricolous,” he said, “their medium is soil.” He went on to explain how each terricolous species needs specific soil chemistry, a certain acidity. 

The same is true of species that depend on other substrates for a home.  Lobaria oregana (Oregon lung lichen), for example, grows best in mid-elevation middle-aged to old-growth forests west of the Cascades. Conifer trees in these forests provide the perfect habitat for this species.

In addition, cyanolichen and tripartite lichen are limited by water and light requirements. They will “fall out at certain elevations,” explained Joe, as the environment is too harsh and dry. They also tend to be found where more light is present. In contrast, they can often be very abundant in riparian areas, where water and light are readily available 

A Peltigera species growing in the soil

Dispersal, Growth, and Reproduction. 

Despite their abundance, lichen still are relatively slow-growing organisms and don’t disperse or reproduce easily. 

Though growth is pretty variable. According to Joe, crustose lichen typically grow only a few mm or ½ cm per year. Foliose or fruticose lichen may grow several cm, depending on conditions. 

When it comes to reproduction and dispersal things don’t get much easier. Sexual reproduction is one possibility and is performed by the fungi. “A high percentage of our lichenized fungi are Ascomycota,” said Joe, when I asked about lichen reproduction. This sort of fungi produces disc-like fruiting structures called apothecia from which asci (spore-containing cells) can be found and lichen spores (ascospores) are released.  

However, sexual reproduction can be challenging for lichenized-fungi that need to not only find a mate but a photobiont. Therefore, asexual reproduction is also common as it has a better chance of success. In the case of asexual reproduction, packaged “dispersal units” are released that contain all the parts of a functioning lichen. Isidia and soredia are the names of these dispersal units, both having some different characters, but function in a similar way. In addition, both isidia and soredia (and patches of soredia called sorelia) are observable, especially with a hand lens, making them also useful for identification. 

What’s not to Lichen?

The snow was turning to a chilling rain when Joe and I decided to turn around. Joe’s wife and son were supposed to be meeting up with us soon. It was at this point that I asked Joe the all-important question—why should I care?  For all their “good looks,” is there something more a person might appreciate about lichen?

“Lichens really tell you about the health of the ecosystem and the health of you and all the animals and all the plants,” Joe responded. “The more diverse the ecosystem and the healthier you will be…. without these organisms we wouldn’t exist.”

This might seem like an extreme view to some, but our dependence on the natural world (as part of the natural world) is well established. And when it comes to the mysterious world of lichen, the value that they provide is as diverse as the species themselves. 

The trail (or road) as we turned to head back down.

Fix it

One of the many benefits of lichen Joe shared with me early in the hike is their “important role in the nitrogen cycle.” The cyanobacteria in cyanolichen (including tripartite lichen) are capable of chemically changing nitrogen gas in the air (N2) into a form that is biologically available to plants and algae. This process is known as nitrogen fixation. Then as lichen fall to the ground and decompose, the nitrogen stored in their tissues becomes available to other life forms that need it.

In certain environments, this input of nitrogen may prove to be significant. Though more research is needed to better understand the extent of their contribution. 

Lichen also plays a role in the hydrological cycle, and other mineral cycles, intercepting and storing water and atmospheric inputs of various nutrients—like with nitrogen, providing a catch basin and distribution system for these inputs. 

Food and Fiber

Lichens also provide food and fiber to other living things. Deer, elk, and caribou feed on lichen, depending on it for winter forage. Squirrels, chipmunks, mice, and bats also take advantage of lichen for food and nesting material. Joe specifically mentioned flying squirrels, which rely on the lichen Bryoria sp. for food and nesting material during certain parts of the year. Flying squirrels are a major food staple for Spotted Owls; without lichen, the entire food chain would collapse.

Biological Indicators

Then there is their importance as biological indicators. The presence, and more often the absence, of biological indicator organisms, acts as an alarm system for environmental change. Like a canary in a coal mine, sensitive lichen will die off in areas where there is too much air pollution, while others may move in. Similar responses may occur with changes in precipitation and temperatures. Therefore the species composition in a location can tell you a lot about the health of the ecosystem and the stressors it may be facing. 

Understanding lichen community composition and tracking it over time has been another large part of Joe’s work with the Forest Service. With over 3,000 plots scattered throughout forest service land, The National Forest Inventory and Analysis (FIA) Program tracks a plethora of forest-related data, including lichen community data. 

Using this data, scientists are able to see which lichen species are present in various climate and air quality conditions.  One of the startling patterns that emerged from this research is the changes in species composition as you move toward urban centers. “ A large number of species are dropping out,” said Joe, due to human impacts like poor air quality.

Changing Climate

In addition, Joe hopes that by observing changes in lichen communities, we might also be able to gain a better understanding of climate change. We know that lichens are sensitive to climate conditions, so it is likely that they will respond to climate change, especially in sensitive alpine and subalpine environments. 

The arctic is already seeing a decline in lichen species as a result of climate change, so the question is how will this translate into other ecosystems? “Timelines for noticing the rate of change is gradual.” explained Joe, but “we have over 30 years of FIA plot community data” to work with.

Lichens to Know

As we continued our chat about climate change, Joe and I ventured our way back down the hill, eventually reaching a bridged creek crossing and an abundance of cyanolichen. 

Bridge Crossing

Joe held up a beautiful, deep brown colored lichen with ridges that ran along the upper cortex that reminded me of a river system and its tributary. Joe told me that the eruptions were soralia, soredia (asexual propogules) yielding structures and that the lichen was a Lobaria anomala (netted lungwort lichen). 

Lobaria anomala

Beautiful, bright green Lobaria pulmonaria (lungwort or lettuce lichen) littered the area, coating the bark of Ash and other riparian trees. And long strands of Usnea longissima (beard lichen) draped across the branches like a garland; a species common to the area, but rare globally. We even saw Joe’s “baby”—Sticta gretae sp. nov. growing on a tree branch.

With so many lichen around, I asked Joe for a shortlist of “ones to know” for the region. He suggested the following based on their abundance and distinctiveness: Platismatia glauca (ragged lichen), Parmelia sulcata (wax paper lichen), any Lobarias, and Caldonias. And almost instinctively, he continued to point each beauty in the vicinity out. 

A tree festooned in Usnea longissima

Cooperation 

At this point, Joe and I caught up with his wife, Elisa and their young son. We were only a few minutes from the trailhead, but we took our time getting back, chatting about life and lichen along the way. 

It was fun to watch the happy family together, harmoniously moving through life together as we walked along the trail. 

It also reminded me of what interested me in lichen in the first place—the symbiotic relationship. The world of lichen is “rooted” (while not having actual roots) in cooperation; in give-and-take; in shared goals.

This sort of relationship might seem baffling at first—isn’t nature a battlefield? A competition with winners and losers?

Perhaps, or perhaps this view is too narrow. Lichen remind me that cooperation is natural—as natural as a family walking together through a lichen-filled forest.

What do you think? What’s not to lichen?

Joseph R Di Meglio is a Mycologist for MICROTERRA Analytical and Pathology Laboratories llc. and Molecular Lichenologist at Oregon State University. He also contracts with the U.S. Forest Service on lichen related projects. Joe studied mycology and lichenology at Oregon State University.

Hike with an Entomologist

I met up with Chris on a drizzly cold morning in November at the top of Lewisburg Saddle in Corvallis before sunrise. Moisture hung in the air and a veil of darkness shadowed our view into the forest. At this point, you might be thinking—why? What possible purpose might Chris and I have for hitting the trail so early?  

The answer is simple—to look for beetles, of course. 

The Hike

  • Trailhead: Lewisburg Saddle Trailhead
  • Distance: Variable (our hike was probably 2-3 miles)
  • Elevation Gain: Variable, but plenty of low elevation gain options
  • Details: Popular trailhead with plenty of parking. Information kiosk at the trailhead. Several trails and logging roads to hike in the area.

A Beetle Guy

An entomologist and curator for Oregon State University insect collection for the last 15 years, Chris is passionate about insects and their taxonomic relationships. In particular, he is a beetle guy.  When I asked: why beetles? His reply was simple: self-preservation. Chris started his insect collection at a young age but soon became overwhelmed by the sheer volume of insects on the planet. There are a lot of insects out there numbering in the 100,00s of thousands to millions. So to simplify things he decided just to focus on beetles. Of course with nearly 400,000 known beetle species on the planet, it is easy to see how his passion morphed into a life-long career. 

Now, net in one hand and the leash of his pupMaera—an Italian truffle dog he hopes to train to detect beetles—in the other, Chris was ready for the hunt! I, on the other hand, a wide-eyed amateur, had no idea what I was doing. 

Chris Marshall ready to catch some rain beetles.

An Enigma

It was 6:30 a.m. and still dark when we entered the forest. Headlamps on, we crunched down a gravel road, working our way deeper into the woods. Ears and eyes open. Our objective: rain beetles. 

Rain beetles (Pleocomidae or Pleocoma) are a unique family of beetles that are only known in Oregon, California, and Washington. When first discovered, rain beetles were considered an enigma. “A chimera of characters,” as Chris put it. They didn’t fit into any known groups of beetles. Brown bodied, hairy around the edges, with a dark, flathead and uniquely forked antenna—rain beetles are standouts in the beetle world. So at the time, scientists lumped them into their own family and that is where they still stand today. 

In the Dark

Rain beetles’ underground lifestyle only adds to its mystique. Rain beetles spend a majority of their lives underground as larvae, feeding off the young roots of trees and other forest vegetation. They dig deep into the soil’s root zone to forage. A fact discovered when a group of graduate students attempted to unearth the secrets of the rain beetle by digging them up and following their trails. “Six feet, eight feet, even 12 feet down” they had to dig to reach the beetles, said Chris. It was not an easy task with little reward.

Surprisingly, rain beetles live a long time underground before they emerge, and the number of years is strikingly variable. “It appears that they can go through seven to fourteen years before they pupate into an adult,” said Chris. “They need at least seven instars,” or seven developmental stages between molts. Then they might turn into an adult or they might not.

“This is not the standard model you are taught in school,” explained Chris— that the stages are fixed (3-5 molts is typical for winged-insects). However, “It does appear that it is more common than we thought.”  Chris told me examples of wood-boring beetles that emerged from antique furniture! Much to the chagrin of the owners, I’m sure.   

End of Days

But Chris and I weren’t out at the crack of dawn to dig up grubs! This was the rain beetles’ mating season. We were here to find full-grown adults. 

“Usually I hear them before I see them,” said Chris as we scanned the area. “They buzz like a bumblebee.” 

Both male and female rain beetles “tunnel up when the soil is wet from rain,” explained Chris. The males take flight to search for a female partner. While the females stay near the hole they dug up, and release pheromones to attract mates. Once they have mated, the female returns to her hole lays her eggs (over a couple weeks) and eventually dies. While the male will most likely be lost to predation within a day or two. Neither have working mouthparts as adults, to their days are numbered. 

Strange to think about it, but these are the end of days for these beetles. Seven years under the earth, a whole life. Then a brief passage into the light—a day or two. Then death. 

How to (try) and Catch a Beetle

“Males are easy to catch,” Chris explained. They sort of bumble around until they catch a whiff of something attractive.  “They are not great fliers,” said Chris. You can scoop them up in a net with little effort. 

However, females are much more difficult to find and catch. Females can’t fly and rarely leave their burrow. Instead, we would need to spot a swarm of males if we were to have any hope of seeing a female. “A bunch of males fighting for a female—” typical male behavior in the animal kingdom.   

However, dawn was approaching and so far we hadn’t heard or seen one male. Let alone a swarm of them. 

A Geographic Mystery

As the day began to grey with pale morning light, we continued our search following the gravel road downslope. Despite our lack of success, our spirits were high and there was still a lot to talk about.  

Chris’ research on rain beetles has mainly been focused on understanding their geographic range—where they are and how they got there. 

“The species can only expand its range at the speed at which larvae can move around,” explained Chris—which is not very fast. And they are also limited by barriers, like rivers and mountain ranges. So the question is: “How did they become distributed on the west coast? This is the heart of my interest,” said Chris. 

Islands of Beetles

Our “trail” lined with great rain beetle habitat

Chris has been deeply involved in collecting and studying the genetics of different populations of rain beetles in order to begin to piece together their story.  

“Historically they were one big species,” said Chris. Or at least that is where the evidence seems to suggest so far.  As physical barriers have formed and shifted, the “one big species” has been split into many—a process known as vicariance. “They (the beetles) were there first, and the river or mountain range came second,” explained Chris. When considering changes over time in physical geography, “there is a nice pattern,” to the distribution of species of rain beetles. 

New Species

Recently, Chris put in the painstaking work of describing a new species of rain beetle on the north side of the Columbia River that separates Oregon and Washington. An impressive accomplishment marked by many challenges. Chris hopes to add the new species to a chapter he is writing on rain beetles for an upcoming edition of a general beetle reference book.

Continuing with this work—Chris is interested in visiting many of the known Oregon populations across the state and getting material. Many of them haven’t been sampled since the 1950s!

Flooded

The population of rain beetles we were searching for is one of many isolated “island” populations found in the Willamette Valley. We were hiking at around 800 to 900 feet above sea level for a reason. Rain beetles are “not found on the Willamette Valley floor, only at higher elevations,” said Chris. Why? Chris’ hypothesis is that the Missoula Floods may have wiped out any lower elevation populations and separated the upper elevation populations from each other.  

The Missoula Floods were a series of massive ice-age floods that scoured the landscape of eastern Washington, ripped through the Columbia River Gorge, and carried water and sediment into Oregon’s Willamette Valley as far as Eugene. Water levels in the valley are estimated around 400 feet in Portland and 350 feet in Eugene—plenty high enough to drown out the beetles. 

“Did you know they were proposed to be the state insect?” asked Chis. I hadn’t.  They lost out to the Oregon Swallowtail—a butterfly. “Not a species, a subspecies,” said Chris. Apparently, colorful, charismatic butterflies are also enough to drown out elusive brown beetles.  

Predawn

At this point, the sky had changed from a dark grey to a lighter grey. Still too early to see clearly. “We are not giving up,” exclaimed Chris, “they fly as late as 10 a.m.” 

Rain beetles are crepuscular—meaning they are active at dawn or sunset. “Their eyes have cones and sensors that are good at seeing at low light levels,” explained Chris. This is an advantage to the beetles and a disadvantage to the beetles’ predators. “Most vertebrate hunters are nocturnal or diurnal,” said Chris. Meaning they either see really good at night or during the day.  Flying at this time means rain beetles are less likely to be eaten before getting a chance to mate. 

Flying at this time also means rain beetles are less likely to be seen by us. Wait, I think I saw something buzz by and fly up into the forest. I must be seeing things.

Ghost Beetles

After about 40 minutes of hiking, Chris and I decided to turn around to see if we might catch some beetles on the way back.

Then, before long, we see it! A ghostly creature bumbles by, greyed by the paleness of the day. Chris springs into action but is too late. We see it again, but still no luck. We even try climbing up the steep embankment of the road cut to try and catch one, but the ground is slick and the route is steep. I slide down on feet and hands, turning the palms of my hands a clayey-brown. 

Chris assures me. There will be others. 

We keep to the area pacing a bit, catching glimpses of “ghost beetles,” before continuing our route upwards in the direction of our cars.  

Describing Science

As we walked, Chris shared with me some of his concerns about the direction science is headed. “In the last 50 years, science has been reduced to this experimentalism,” he said. It has become mostly about repeating controlled experiments. Experiments are all well and good, but by limiting our focus we “cut out a hugely important part of science,” Chris said. 

“Science has a descriptive component,” he explained, “that is heavily embedded in the exploration and discovery phase.” He went on: “At the end of the day, I am trying to show you that this thing exists. If I can show you one, then you will know.” 

This is also why pinned specimens and research museums are so important.  “Museum specimens…provide empirical observation. They serve as an archive for scientists to describe the northwest, country, and even the world.”  They show us what is out there. 

In addition, “tests can be run on specimens,” to determine taxonomic relationships. This is what Chris’ work is all about. Describing new species. “So we can say that this is X and this is Y.”

“There is so much to discover!” he exclaimed. As Chris puts it—descriptive science is “how we know what is on planet Earth.” And—how can we care if we don’t know?

Associations 

It was now getting to be much lighter out. We had “seen” several ghostly rain beetles, but were yet to catch one.

Out of the blue, Chris B-lined it toward a cluster of orange fungi on the side of the road—possibly toxic“Jack-o-lantern fungi.” He picked it up and looked under it to see if he could spot any beetles. He explained that many beetles live in and on mushrooms. Many are “fungivores”— consuming mushrooms or other parts of the fungi. 

This got Chris reminiscing. “Thought I would study beetle-fungi coevolution,” he said. “But it was tough to get funding.” Then it was ants and beetles. But the fieldwork was hard and he didn’t want to be an “ant biologist.” At the same time, he was fascinated by associations between beetles and other living things. So when he found a relatively abundant beetle was covered with mites— he changed his project again. 

“What did you learn?” I asked

“I learned PhDs take a long time,” was Chris’ curt response. “You can quickly bite off more than you can chew. You have to learn to spit some out.” Ultimately, though he completed the project, he didn’t end up getting into the mite-beetle specifics as he had hoped. Does he regret it? Yes and no. “I have a full life of other things I like to do too,” he said. And that is also important.

Orange mushrooms, possibly Jack-o-Lanterns.

How to (finally) Catch a Beetle 

Just then, something caught Chris’ eye! One of our ghostly friends came flying down out of the forest. Only, this time Chris was quick to action, swooping his net in a well-choreographed flip of the wrist and he caught it! 

Chris pulled the rain beetle from his net and held it out in his hand.  “There he is in all his glory,” he said. 

I looked down at this scruffy, male beetle with his shiny nearly-black exoskeleton fringed with golden brown fur. He was both glamor and gruff. At least that is my take; granted I am no entomologist. 

Chris passed him off for me to hold. But in my hands he squirmed so much that I soon passed him back. 

It was a fleeting moment for both of us. 

Our first catch! Chris holding a rain beetle.

The Beetles Go On

After catching one, Chris ended up scooping up a couple more rain beetles before we made it back to the trailhead. We walked and talked and caught beetles—Chris sharing funny anecdotes, and thoughts on career paths not taken. 

Toward the end of the hike, I asked Chris what else he could tell me about finding beetles. He responded with a litany of places to look: under rocks and logs, in flowers, under driftwood in dunes, on plants, and swimming in water. “There are a ton of different ones,” he said. “A lot are small too,” he explained, and “need a hand lens to see them in detail.” 

“We have an incredible native fauna,” said Chris. And to experience it all requires incredible effort—learning habitats, timing, etc. of the beetles. In short, to “be an insect collector,” you need to know your insects. 

Individual Style 

Chris is all about studying the individual and species levels in biology. He was firm on that point. “I am not an ecologist,” he stated frankly early on in the hike. Not that he doesn’t enjoy ecology, it is just that Chris uses a different scientific lens.

Chris used the example of a tree in a forest to explain his perspective. “An ecologist has to reduce the tree to a primary source of photosynthesis… a carbon sink.” At the same time, a reductionist might see the tree as just a “physiological unit of cells.” This is not complete either.

While both these perspectives are important and valid, they miss another equally important and valid way of studying nature. “Sitting in between those two things. The thing that links those things—the individual and species level.”

“That’s an individual,” said Chris as he pointed to one of the trees in the forest. And that individual comes from a pedigree of individuals… That’s the realm of systematics and taxonomists.”

Caught

Chris extended his thinking to human societies as well. If we are “so focused on how the bigger system works,” he argued “what does one persons life matter?” 

As a systems thinker myself, I often think a lot about big picture ideas, so I was refreshed by Chris’ ideas. To consider the individual—whether it be a human or another life form—makes a lot of sense.

Every human, tree, or beetle has its own unique biology and its own unique life history. And as helpful as it can be to see the forest through the trees—generalizing can also be a dangerous game. You lose out on important details. Diversity is lost. Judgments and decisions are made without all the information. And that has the potential to be catastrophic. 

By the end of our hike, Chris had caught three beetles. But I caught something as well, perhaps even more valuable—a fresh perspective.

Dr. Chris Marshall is the  curator for the Oregon State Arthropod Collection.  He earned his bachelor’s degree at Reed College before studying entomology as a graduate student at Cornell University.

Hike with a Land Conservationist

Sunset at Whychus Canyon Preserve.

In my experience, relationships with places are not all that different from relationships with people. You have to spend time with a place to get to know it. Ask it questions. Become familiar with its moods and seasons. Learn what makes it tick. Before long, an intimacy may develop and you may even find yourself saying the L-word—love. It takes time. Sure there are those love-at-first-sight moments—but those are fleeting. A deep relationship to the land is more than a few moments on a clifftop view watching the sunset. 

When it comes to Whychus Canyon Preserve, few people have a deeper relationship with the land as Sarah Mowry. As a staff member at the Deschutes Land Trust for the last 15 years, Sarah has been with the property since it was first established in 2010, and in 2014 when an additional 480 acres were added. With that in mind, I met Sarah at the Whychus Canyon Preserve trailhead on a cool autumn afternoon to explore the place for the first time.

Sarah Mowry making her way along the trail.

The Hike

  • Trailhead: Whychus Canyon Preserve Trailhead
  • Distance: about 4 miles
  • Elevation Gain: approximately 500 feet
  • Details: Directions and details are found at the Deschutes Land Trust website.

Welcome to Whychus

Before hitting the trail, Sarah gave me a bit of background on the Whychus Watershed. Whychus Creek is a glacier-fed stream. The creek begins up near the Three Sisters, tears downhill until it reaches Alders Springs and its confluence with the Deschutes River. 

Most of the river’s path is marked by deep canyons, but there are some places where the land opens up and meadow habitat is possible. According to Sarah, Whychus Canyon Preserve, the property we were about to explore, has some of the best meadow habitats.  These meadows are “biological hotspots,” Sarah explained.

In addition, Whychus Canyon Preserve provides habitat features for Chinook salmon and steelhead, which are being reintroduced into the Deschutes River Basin. “There has been a huge collaborative effort to bring them back led by the Confederated Tribes of Warm Springs and Portland General Electric,” explained Sarah.

The preserve also provides migratory routes for terrestrial species, like deer and elk, as they move down into their winter range. “The habitat connection the Preserve provides to adjacent public land is huge,” said Sarah,. And at 930 acres, the Whychus Canyon Preserve extended the habitat substantially.

Plus, Whychus Canyon Preserve has an extensive trail system with 7 miles of established trail for people to explore.

All this to say, Whychus Canyon Preserve has a lot going for it. 

Trailhead kiosk provides background information about Whychus Canyon Preserve

Restore

After several minutes discussing the property, Sarah and I realized we better hit the trail if we were going to finish our hike before sunset. We had decided on a 4-mile loop down to the river and we immediately began our descent.

As we tripped downhill past dried bunchgrass and sagebrush and past juniper and pine, Sarah told me about the Land Trust’s forest restoration work. She explained that when the Land Trust first acquired the property, the forested canyon was thick with small  juniper and pine. So in order to restore the land, some of the trees were thinned out by hand.

Restore. Restore is a tricky word. It means to return to its former state. But how far back do you go?  Can cutting down trees really be considered restorative?

The short answer is—it depends! Restoration work, as Sarah explained it, all depends on the location, local ecology, and the project goals. For the Whychus Canyon Preserve, cutting down a few trees made sense. It helped with fire protection and opened up the forest for larger pines and junipers while promoting healthier habitat for all sorts of other plants and animals.

Jumpstart

One of the Land Trust’s goals for  Whychus Canyon Preserve, is to “restore a natural functioning system,” said Sarah. And, sometimes, a hands off approach won’t get you there. Past hands have already had an impact, so expecting nature to bring it back just isn’t going to happen. Healing the relationship between the land and people requires time and work. “You need to jumpstart the system,” said Sarah, “Give it a leg up so it can get itself back to a healthy place.”

Of course, the way you do so can be tricky.  For example, in Whychus Creek at nearby Camp Polk Meadow Preserve, another Land Trust restoration site, the stream was restored by digging out historic channels and adding curves and other features for habitat complexity. Fast forward four years and the Land Trust is working with partners to restore Whychus Creek at Whychus Canyon Preserve using more process-oriented methods. “We are learning things all of the time,” Sarah said, “The kind of restoration work we were doing now has evolved from what we did 10 years ago.”

Canyon Bottom

Before long, Sarah and I had made it down to the bottom of the canyon and the crystal clear waters of Whychus Creek. Trees and shrubs line much of its banks, as it continues cutting its way down deepening the canyon.

“We haven’t done any stream restoration in this part of the Preserve yet,” said Sarah. Eventually, she explained, a detailed plan is currently being created that will lay out everything—stream structure, plantings, habitat features, etc. Thousands of native plants will be brought in to fill in the gaps. And logs—lots and lots of whole trees are needed. “Large woody debris,” as it is often called, provides cover for fish and aids in the development of stream habitat diversity.

Whychus Creek at the canyon bottom.

Free

As we hiked along the bottomlands, Sarah pointed to areas where strips of land had clearly been raised adjacent to the creek; probably dug by the Army Corps of Engineers with good intent to reduce flooding.  Instead, these berms disconnected the stream channel from its floodplain, limiting the ability of Whychus Creek to spread out.  Thus giving the  creek access to its floodplain will also be an important part of the restoration plan.

However, this doesn’t mean the creek will simply be rechanneled—directed by the will of people. Instead, a process-based restoration is being implemented throughout the Whychus Canyon Preserve. With this sort of plan, the Whychus Creek will be free to find its own path, or paths, as it were.

The Land Trust has already begun using this sort of methodology on the northernmost mile of recent creek restoration efforts at the Preserve. Left to find its own path, Whychus Creek has created several new channels and water is beginning to saturate the surrounding landscape. In fact, some of the pines in the floodplain are dying off because the soils are now too wet to support them. “It’s a little hard to watch,” said Sarah, but it’s all part of the process. Those trees will become homes for other animals as snags or provide cover for fish.”

The newly wet floodplain also meant a different planting plan for the restoration. When you let the creek choose where it will go, you can’t choose where to put the water-loving plants or the plants that prefer dryer conditions, so you plant a little of everything everywhere, explained Sarah.

Berms along Whychus Creek disconnect the stream channel from its floodplain.

Healing

Restoration isn’t hands-off, but all hands on deck. It is work. The land comes with “baggage” from past human relationships—sometimes scars. Restoration is providing the opportunity for renewal, a starting point. Then knowing when to back off and let nature heal itself.

Watching the land heal is a huge perk of Sarah’s 15 years with the trust. “It is awesome because I get to see the positive changes we can make over time.”

Land Trust

Earlier on during our hike, Sarah pointed out several houses built along the rimrock on the opposite side of Whychus Creek from the land trust’s property. Now, as we began to make our ascent back up the canyon, more houses came into view perched above us.

“Development is a challenge to nature,” explained Sarah.  “A lot of good habitat is on private land. What can you do when it is covered over?”

That is why land trusts are an important part of the land conservation equation. Land trusts fill in the gaps where public lands can’t by strategically purchasing lands, or establishing land protection agreements called conservation easements, with a focus on the future.

The Deschutes Land Trust was first established in 1995 for this very reason. Central Oregon was developing at a rapid pace and many community members were concerned about the loss of wild areas and vital habitats. So when a well-beloved parcel of land went up for sale and was threatened with development, community members came together, and the Deschutes Land Trust got started, by protecting their first property—the 63 acre Indian Ford Meadow Preserve just outside of Sisters, OR.

Now the Deschutes Land Trust manages over 17,000 acres of land. 

View across Whychus Creek where houses line the rimrock.

Easements

Nearly  half of the Land Trust’s lands are  protected through land protection agreements called conservation easements. Conservation easements are agreements with landowners to protect or restrict certain activities on their private property in perpetuity. Each agreement is unique to the land and the owner.

Why would a landowner want to put an easement on their property? “Most of our landowners have done it because they have a conservation vision,” said Sarah.

View at the Top

The trail steepened as Sarah and I climbed some rock steps, sweating our way to the top of the canyon and a sweeping view looking out over the preserve. Sarah told me that the trail builder that put in the steps we were climbing was all about “the journey instead of the destination.” But I have to admit, the destination, in this case, was sort of the point. 

Looking out across the canyon, layers of rim rock were imbued with a golden hue, and stately pine trees mixed with juniper dotted the landscape all the way down to the fall-colored leaves of deciduous trees that lined Whychus Creek. In the distance, you could just make out the meadow that Sarah had talked about earlier. This was one of those moments—a beginning—a connection to the land. 

We paused here to take in the scenery and experience the preserve from a different vantage point. Sarah pulled out a map of the area that showed a conceptual rendition of the new stream channels that were forming and reforming as the meadow has been restored with an influx of water. And we talked at some length about restoration monitoring methods and the wonders of lidar imagery.

But the sun was getting lower and we had families to get home to, so we made the difficult decision to continue onward, following the trail along the canyon rim to our cars.

View looking out across the canyon.

Time for Change

As we walked along patches of old growth juniper and sagebrush steppe, Sarah and I talked about the people in our lives and the changes we have been dealing with lately. We discussed the challenges of having kids in distance learning, changing job responsibilities, and just a general sense of loss that life has taken lately.

One of Sarah’s responsibilities as outreach director is to coordinate events that bring people to the preserves to learn more about it. People need to “learn about a place to care about a place,” Sarah explained. And for the time being, these sorts of events are just not possible.

Renewal 

Looking around at the dried out sagebrush and bunchgrass along the path, it is difficult to imagine anything else. But each spring the brown earth is renewed with bright fields of green and colorful spring wildflowers. “Gold stars blanket the floor,” Sarah said in remembrance. The “star” of the show are dime-sized goldfield daisies that bloom in early spring, enveloping the land in a warm yellow profusion of color.

The winter we are facing right now makes it seem like we will never see spring. But barren landscapes can be returned to beauty and function whether through changing season, or, at times, through restoration.

This makes me think, perhaps that is what is really needed—a restoration. To be brought back to an ancient connection with the land, and with the people that inhabit it. We need to turn to one another and turn to the land. Nurture relationships. Listen and learn. It will take work—hard work—and a good deal of patience, but if we can get things moving in the right direction, perhaps nature will kick in and bring us back to something better.

Now that is a change I would love to observe.

Sarah Mowry is the Deschutes Land Trust’s Outreach Director. Sarah has been with the trust for the past 15 years. She has a Bachelor’s Degree in Environmental Studies from Middlebury College and a Masters Degree from the University of Montana.

Hike with a Cell Biologist

View out toward Marys Peak from the top of the hill.

Let’s talk tiny. I mean really tiny. Like, get out your microscope small. I am talking about cells!  You know them, you love them—those little bags of gobbledygook filled with smaller stuff still, with names like endoplasmic reticulum, ribosomes and Golgi bodies. 

Cells are the building blocks of every living thing on the planet. So despite their size, they are kind of a big deal. So big that on a sunny Thursday afternoon I met up Marc Curtis, a cell biologist from Oregon State University, to talk about the small stuff.

The Hike

  • Trailhead: Forestry Club Cabin Trailhead at Peavy Arboretum, Corvallis, OR
  • Distance: 4.8 miles
  • Elevation Gain: approximately 900 feet
  • Details: Plenty of parking at the trailhead. Restrooms available at the arboretum, but not at the trailhead. This hike is part of an extensive trail network throughout the McDonald-Dunn Forest, so there are many options to extend or shorten the hike.

Cell Level Thing

We immediately started hiking uphill from the parking lot along a forest road that leads to Oregon State Universities Forestry Club Cabin, before turning onto the tree-lined trail. As we trudged along several steep sections of the trail, Marc told me about his background. 

Marc was fascinated by cells from a young age. In high school, his father, a Molecular Biologist at the Wistar Institute in Philadelphia, gave Marc a review article on cells and their role in cancer. Marc was hooked. “That just made sense to me, that cell level thing,” he said. “They grow, divide, they talk to each other, they take on different functions and they build the body.”

As an adult, Marc pursued his interest in cells by studying biochemistry at the University of New Hampshire and getting involved in undergraduate research involving cell signaling in the corpus luteum. Later, he moved onto Oregon State University to study cell death, as a graduate student, and eventually cell mutation as a postdoc. 

Grow, Divide, Repeat

Marc has made a career out of thinking about cells, what they do, and how they do it. But cells are so small, invisible to the naked eye. Can we find a way to appreciate cells while on a hike?

A good start is to pay attention to plants. This is where Marc has focused most of his career. 

Unlike humans, plants typically grow throughout their lives—packing on the inches at their tips, as long as conditions allow. Plant growth occurs in the roots and the shoots—where apical meristematic tissue composed of undifferentiated cells grow and divide. Meristematic tissue is also found in the buds and the nodes, the “joints” of a plant. Many different plant tissues and organs can arise from these growth regions, including leaves and flowers.

Marc pointed out the shoot apical meristem on one of the plants we saw along the trail. “It is in there,” he said, directing my eyes to the place where the leaf meets the stem. “That is where you have the undifferentiated cells—the fountain of youth—where new uncommitted cells come from.” 

Marc explained that when these cells divide, they build up from the bottom layer and will eventually differentiate and become tissue, so the cells at the tips can maintain their undifferentiated status. 

Meristematic tissue is found where you see branching on this plant.

Time for a Change of Pace

However, this doesn’t mean these cells remain unaltered throughout the life of the plant.  Because apical meristem divides so prolifically, mutations can accumulate in the meristematic tissue.  Marc studied the process of mutation using meristematic tissue as a postdoc. Mutations are “mistakes” in DNA—a cell’s molecular instruction book—that arise during the replication process or from environmental factors, like UV light. 

Marc was interested in how plant cells are able to bypass mutation so they can continue to grow and divide. Turns out that plant cells are pretty good at this. By maintaining a low level of fidelity during the replication process, cells in the apical meristem can continue their work of supplying the rest of the plant a lifetime supply of—well—cells! This also means there can be 100s of genetic variants in the meristem of one individual plant that can potentially give rise to unique growth forms. Though it seems like this is fairly rare. 

In any event, growth is cellular! So when you see new leaves and flowers emerge in the spring, or look up to the tippy-top of a tree, or notice a new growth pattern in a familiar species—think tiny! Think— grow, divide, repeat! The leaves, flowers, and new growth each year are the result of the microscopic world of cellular division! 

Looking up at the tippy-top of some trees on the trail.

Beauty in Death

In autumn, cells take a turn for the morbid. As Marc and I made our way further up the trail, leaves crunching underfoot, I asked him to explain how cells were involved in the spectacular displays of fall foliage observed during this time of year. 

The process is called autumn senescence—which essentially means a slow, seasonal death. Cells that make up the leaves of deciduous trees start to shut down in the fall in response to changes in daylight hours and temperature. In order to conserve resources, cells “break down chlorophyll and other components,” Marc explained, “leaving carotenoids and other pigments exposed.” Hence, the bright oranges, yellows, and purples. This organized way of dying, allows plants to hold onto difficult to obtain nutrients, like nitrogen, so that later in the spring they can begin to” grow, divide, repeat” once again. 

Bigleaf Maple autumn color (a.k.a. dead/dying cells)

Dead Tissue Eater

However, cells don’t always die in a blaze of colorful glory. Cell death may also be an adaptive defense. Earlier in the hike, Marc talked about his PhD work on a plant that when attacked by a toxic fungus would respond by activating cell death. This might seem like a bad idea at first glance, but by killing off the tissue where the fungi attacked, the plant was able to stave off further damage and prevent the fungus from eating it. You see, this particular fungi was a biotroph, meaning it only consumes living tissue. Dead tissue was entirely unappealing.  

Unfortunately, Marc’s tale has a sad ending. Another fungus came along—a necrotroph, or dead tissue eater—that was able to mimic the biotrophic fungi’s toxin, triggering cell death in the plant. Only in this case, the dead tissue was very appetizing to the fungi. It is a dog eat dog cellular world out there! 

Gall-y

Eventually, Marc and I made it to the top of the Powder House Trail, where the vegetation changed from the Douglas-fir and Big Leaf Maple trees we had spent most of our hike walking through, to a hilltop prairie of grasses and Oaks. Walking amidst the oak trees reminded me of the dozens of oak galls my kiddos and I had spotted amongst the fall foliage on another recent hike in the area. I asked Marc about these odd growths. 

Though he didn’t know a lot of details, Marc did recall that the growth was the result of a parasitic “gall wasp.” The gall wasp will lay its eggs on an oak leaf or twig, usually in the spring. Then, when these eggs hatch into larvae, they release biochemicals that “brainwash” the cells of the leaf or twig into forming a gall. The gall—a growth filled with nutritive cells—envelops the larvae as it forms, protecting and feeding it while it pupates. 

Upon further research, I found that gall shape and structure are unique to each parasite, even between individuals of the same species. And that the growth pattern of galls are so different from normal leaf or branch tissue growth, that “they have been described as new plant organs” with a unique chemical signature. Though it seems there is still more research needed to understand exactly how the larvae control cell growth and division, galls are an impressive example of what I like to call “zombie biology.”

Green Islands

In discussing galls, Marc was reminded of another parasitic relationship that can be observed in the fall—green islands. Green islands are spots on a leaf that will remain green even as the rest of the leaf begins to undergo senescence. Marc explained—the green spots are the result of a fungal infection. “The fungus produces plant hormones called cytokinin which delays senescence.”  This keeps the chlorophyll—the green stuff involved in photosynthesis—from breaking down. Hence the green. This also means the plant’s food factory stays open for business, providing the fungus a continuous supply of sugar. Needless to say, this is a pretty sweet deal for the fungi. 

Green islands on Bigleaf Maple leaves

Wax on, Wax off

As Marc and I meandered along the trail looking for green islands (later we found a few examples), Marc pointed toward a gnarled and twisted tree trunk with peeling red bark straight ahead—a Madrone! 

A personal favorite of Marc’s I walked up to get a closer look. I felt one of the thick, smooth leaves.  “What is going on here?” I asked. “Why do Madrone’s have waxy leaves?” It must be something cellular I thought. 

I was right! “Waxes are secreted by the epidermal cells through the endomembrane system,” Marc replied. The endomembrane system is the machinery in a cell that packages and transports certain molecules, like wax, into the extracellular world. Marc explained, The Golgi probably synthesizes the wax. It is then gathered in vesicles. These vesicles transport the wax to the outer cellular membrane. Here they fuse with the membrane and dump the wax out into the cell wall. And voila—waxy leaves! 

Human bone-building also uses the endomembrane system, Marc shared. Though in this case, the bone cells are spewing out collagen protein—the number one ingredient for bones. Think about it! Our skeletons are made from cells throwing up all day. A fun fact Marc likes to share with his students.    

Waxy leaves on a Madrone tree.

Evolution

Eventually, Marc and I made our descent back towards our cars. On our way down the trail, we chatted about topics ranging from teaching on zoom to plant podcasts. 

Marc was also on the lookout for a liverwort he had been able to identify recently and wanted to see if he could find it again. In addition to several botany courses, Marc also teaches an evolution course at OSU. So with ancient plants on the brain, biological evolution naturally came up as we hiked along. 

Innovation

Like most things, if you haven’t caught on yet, evolution is very much a cellular process. Mutations that occur in reproductive cells, gametes, provide new genetic variation to a population. Meiosis, the development of reproductive cells, does the same by mixing and matching genes so each gamete is unique. 

Of course, there is much more to evolution by natural selection than genetic variation. When passing by a Douglas-fir tree, Marc shared with me his thoughts on the subject. “Bark,” he said, “has a high surface area with all the cracks and creases.” When bark like this evolved it provided opportunities for many other species to evolve on it. “Innovation opens the door for more innovation,” Marc explained. He used the analogy of the internet boom.  Once the internet got started it provided opportunities for online retail, social media, etc. One new idea brought about many more ideas.  Life is similar—a new biological innovation can open up ecological space for new species to emerge. 

Douglas-fir bark with deep ridges.

Listen to a Liverwort

Toward the end of our hike, Marc finally found his liverwort. He pointed out how the “leaf” structure and arrangement was different from that of a moss or any other plant. A difference that Marc had only recently developed an eye for, and I had never considered. In fact, I was pretty sure I have been mistaking liverwort for some other group, like a lichen or moss, my entire life. 

Marc’s liverwort! Check out those “leaves!”

Back at home, I kept thinking about Marc’s liverwort and his thoughts on innovation. And maybe because the world seems so divided, or perhaps it is a personal crisis of faith in mankind, but I can’t help but feel like there is some sort of deeper message here. 

Throughout our hike, the concept of curiosity being essential to scientific work kept coming up. But I think it goes beyond science. We all need to be curious. Open our eyes to the liverworts of the world—not lump them together with lichen or a moss—assuming they are all the same. We need an evolution of the mind! And just like the bark of a tree, innovative ideas will open the mind to more innovative ideas. 

Cells are amazing in their simple mantra—grow, divide, repeat. But it is mutation, change, or innovation—that keeps things moving forward— evolving. 

So, let “new variations” or ideas sink in and become part of your mental framework. If nothing else, you may finally learn to identify a liverwort.

Marc Curtis is an instructor at Oregon State University in the Department of Botany and Plant Pathology. Marc has a Bachelor’s degree from the University of New Hampshire in Biochemistry and PhD from Oregon State University where he studied mutation in plant cells.

Hike with a Wildlife Biologist at Finley National Wildlife Refuge

Looking out at the restored Oak Savannah on the Mill Hill Trail.

When Euro-American settlers arrived in the Willamette Valley of Oregon in the mid-nineteenth century, they encountered a landscape far different from what you see there today.  Historical accounts describe open fields of tall grasses and wildflowers with a few oak trees interspersed.  A biodiverse paradise for songbirds, butterflies, and other species that rely on an open system. Maintained by the indigenous people, the Kalapuya, for centuries—the expansive landscape must have been appealing to many early settlers as well.

Flash forward to modern times—and the Willamette Valley is now a complex of agricultural fields and urban and suburban environs. Most of Oregon’s major population centers, like Portland and Eugene, as well as the state capital, Salem, are in the Willamette Valley. Currently, more Oregonians live and work in the Willamette Valley than in any other part of the state and over 170 crops are grown there. Development has completely altered the landscape. Only about 1% of the Oak Savanna habitat that was once prominent in the area remains. Wetlands, riparian areas, and oak woodlands have all suffered major losses in the Willamette Valley.  

A Few Hold Outs

However, there are still a few holdouts and a lot of effort to maintain and restore what remains of these important habitats. One of the holdouts is Finley National Wildlife Refuge, just a few miles south of Corvallis, OR. And one of the people putting forth the effort to maintain and restore is Nate Richardson, a wildlife biologist for the U.S. Fish and Wildlife Service (USFWS). 

I met up with Nate at Finely to hike the Mill Hill Loop and chat about Willamette Valley habitats and the challenges of maintaining and restoring them. 

Nate Richardson on the trail.

The Hike

  • Trailhead: Mill Hill Trailhead (or Refuge Headquarters)
  • Distance: approximately 3 miles
  • Elevation Gain: 300 feet
  • Details: This is an easy hike. Most people park at the overlook for the display pond. You can also start at headquarters where there is ample parking. However, when this was written, the lot was closed to visitors. There are bathroom facilities within the refuge.

Into the Woods

Nate and I started out on the trail hiking through oak woodlands. Nate pointed out the structure of the forest. There were a lot of smaller, younger oak trees with a few larger oaks and an understory of shrubs—typical of this habitat type. Oak woodland is a priority habitat in Oregon and provides for many species of concern in Oregon, such as the acorn woodpecker and white-breasted nuthatch. Though we didn’t see any during our hike, the Mill Hill Loop is an excellent spot to look for woodpeckers, said Nate. 

A Tale of Two Trees

However, despite the many functioning aspects of the oak woodland we were hiking through, it didn’t take long before we spotted one of its biggest threats—conifers.  Nate explained how Douglas-fir trees grow much quicker than Oregon’s oak species, like the Oregon white oak. Because of this difference, when Douglas-fir inundate an oak woodland, they tend to outcompete the oak by shading them out.

As we continued down the trail, we could see several examples of Douglas-fir trees in direct competition with the oak.  We even saw a few dead oaks as a result. 

A Douglas-fir tree in competition with an Oregon white oak.

Managing with Fire

In the past, conifer and shrub species were kept in check by the indigenous people of the Willamette Valley—the Kalapuya. For hundreds of years, the Kalapuya used fire to maintain the open habitats of the Willamette Valley for forage and hunting. Without fire, woody species have encroached further and further into the foothills of the Willamette Valley, such that much of the oak habitats, especially upland prairie, has been all but eliminated. 

Mixed Up

As we hiked, further along, the suite of species gradually changed from oak to a mixed forest, to one dominated by conifers. With a much denser overstory of Douglas-fir, our path became almost completely shaded, and oak became non-existent. Instead, big leaf maple made up the deciduous canopy with many shade-loving (or at least shade-tolerant) plants in the understory. Sword fern and snowberry were two species I spotted. 

In general, conifer forests like the one we were hiking through are very common in Oregon and becoming more common in the Willamette Valley.  But that doesn’t mean they don’t have value. Mixed/conifer forests provide habitat for many species, such as black-tailed deer and Swainson’s thrush. Nate told me that even Grey Jay (a species I usually associate with higher-elevation conifer forests) will occupy the site in the winter.  

A Douglas-fir forest with big leaf maples and sword fern.

Everyone Likes Ducks

At one point, the dense trees opened up and a brightly lit clearing of grasses and willows came into view through the thicket. A wetland! Wetlands are areas that tend to retain moisture for a large part of the year.  These soggy bottomlands also provide critical habitat for many species, like wood ducks and beavers that frequent the area. Wetlands are also a target habitat for restoration and protection, as many have been drained for other uses. 

Though Nate works for USFWS, he spends most of his time working on restoration projects that are on private lands; many of which are wetland projects. “People like ducks,” said Nate, so they tend to be the focus of these projects. In fact, Finley was originally established to provide nesting habitat for waterfowl. In this case, the Dusky Canada Goose lost a great deal of habitat from land subsidence following an earthquake. Of course, now the refuge takes a multi-species approach to management, focusing on restoring habitat for many species. Coincidentally, even if a wetland is established in the name of duck, many more species will benefit, including humans. Wetlands provide tons of ecosystem services! Wetlands are amazing water regulators and filters, for example. 

Pure Gold

Then, at long last, we arrived! An upland prairie! One of Nate’s favorite habitats in the refuge. According to Nate, upland prairie habitat is really important to many species of concern in Oregon, like the Oregon Vesper Sparrow and the Streaked Horned Lark.  Plant species like Kincaid’s Lupine, Nelson’s Checkermallow, Bradshaw’s Lomatium, and Willamette Daisy, are also rare but can be seen here. Plus oak savanna habitat is typically the only place in Oregon’s Willamette Valley you are likely to find the state bird—the Western Meadowlark.  

Dominated by grasses, like Roemer’s Fescue, and herbaceous wildflowers, like Oregon sunshine, oak savanna is a colorful Smörgåsbord in early spring. The diversity of grasses and forbs in oak savanna habitat also attracts a diversity of insects, which in turn attracts a lot of birds. Essentially, the entire system is driven by the right mix of native prairie vegetation. 

The view looking out at the restored oak savanna.

Time and Money

However, when it comes to restoration, getting that right mix can be a huge undertaking. Nate explained that the open prairie we were looking at used to be a lot of douglas-fir trees and hardly any grasses.  It took a lot of work, time, and expense to bring it back to a near-native state. Cutting down trees, mowing, and burning, as well as replanting native species, are all part of the restoration process.  A process that doesn’t really have an end, as continued mowing, burning and plantings are often needed to maintain the habitat. For example, golden paintbrush, a plant that was once extirpated from the state, needs regular burning to be maintained. Also, as a hemiparasitic species, golden paintbrush benefits from associations with other native species, like Oregon sunshine. 

Visit this spot in spring to see golden paintbrush in bloom, along with a whole host of other wildflower species.

Is it enough?

As you can see, restoration work can get pretty complicated. Research into understanding what species do best, and in what conditions, is another important component of restoration work. However, there is also the question of just how much to restore. Do you want an oak savanna that is 90% native, or will 60% do? Nate talked about the challenges around this sort of decision-making. If you can get an ecosystem 60% restored for a lot less cost and effort, maybe that is enough to restore the ecological function of the prairie. And if that is the case, shouldn’t we stop there? He didn’t have an answer. Nor do I. But these are important considerations for any restoration and/or management plant. When do we let nature do its thing? 

Connect the Dots

Another perhaps even bigger challenge when you are trying to restore an ecosystem that is about 99% lost is connectivity.  Nate explained—Many species need large tracts of land and the ability to migrate between and through the landscape in order to obtain desirable population densities.  Population density is simply the number of individuals in a population that live in a given area. “Song birds really need it,” said Nate, “small places are great, but song birds need more.”  When there is a large enough tract of land that is not segmented, the densities of birds, like the Western Meadowlark, are substantially increased. 

The question is how do we create connectivity? We don’t want the “traditional corridor of trees,” said. Though there is no simple answer, Nate does hope to improve conditions where he can. As mentioned earlier, a lot of the restoration work Nate does is off the refuge property, often in locations adjacent to a refuge. Helping landowners establish habitat on their property can expand the land area that supports species. Then, as Nate described, others see what is happening and want to get involved “and things spiderweb out.” 

A Rare Sight

One of the rarest habitats on the refuge (and in Oregon) is the wet prairie. Due to their location, Occurring in lowlands, especially floodplains, many wet prairies have been converted to agricultural land. Wet prairies are also different from upland prairies as they retain water for a portion of the year making them ideal for plants that like to occasionally get their feet wet. Water-tolerant grasses, like tufted hair grass, sedges, and wildflowers dominate wet prairies.  Unfortunately, the wet prairie is also rare on the Mill Hill Loop—we didn’t see any. 

According to Nate, you have to go to the Prairie Overlook to see wet prairie habitat. Besides being wet prairie, the land adjacent to the overlook is designated a Research Natural Area, set aside for education and research. It is also really special because it is one of the few places in the Willamette Valley that was never tilled. 

Looking out at the wet prairie from the Prairie Overlook.

Important Matters

As we finished the loop and made our way back to park headquarters, I asked Nate why people should care about protecting and restoring wildlife habitat. For Nate, it is all about awe—only in these places can you see a rare blue butterfly or hear a woodpecker cackle, or watch ten thousand geese take flight off the marsh. Experiences like these can inspire people to care about and appreciate the ecosystems around them.

Awe and Inspiration

We were nearing the golden hour as Nate and I parted ways both to our respective homes. But before I left, I pulled over at the Prairie Overlook to take a peek at the wet prairie Nate had mentioned earlier. With my mood light and the sun setting low in the sky over an expansive golden landscape, I really did feel a sense of awe and appreciation. It is amazing what a little bit of nature (and science) can do for the human spirit.

Nate Richardson has worked as a wildlife biologist for the USFWS for 12 years restoring native habitat for the Partners for fish and wildlife program. He got his BS in wildlife science at Oregon State in  2004 focusing on avian conservation and management. In his free time, he enjoys hiking, climbing, fishing, and spending time with his 9-year-old son. 

Hike with a Paleontologist

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

How would you like to travel back in time?

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

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

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

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

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

The Hike

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

Rock Records

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

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

Looking down the trail into the badlands.

Condon 

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

As Nick relayed the story:

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

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

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

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

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

Fossil Land

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

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

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

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

Floodplain Deposits

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

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

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

Colorful Rocks

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

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

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

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

A Layer of Ignimbrite 

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

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

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

Hidden in the Rocks

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

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

Collecting Fossils

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

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

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

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

Past Life

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

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

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

Herbivores

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

Carnivores

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

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

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

A Word On Plant Fossils

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

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

Replicate

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

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

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

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

Replicate

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

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

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

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

Small Things

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

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

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

A river channel deposit of sandstone.

The Species Problem

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

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

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

The Specialist Problem

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

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

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

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

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

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

Amphitheater 

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

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

The amphitheater is the end of the trail.

Storytelling

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

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

I couldn’t agree more, Nick.

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

Hike with a Botanist

View of Iron Mountain from Cone Peak.

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

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

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

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

Linda Hardison on the trail.

The Hike

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

Plants Rule

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

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

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

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

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

Plants are everywhere

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

A Flora 

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

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

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

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

Ecoregions 

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

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

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

What’s in a name?

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

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

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

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

False Solomon’s seal (Maianthemum racemosum)

Not so Common

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

Taxonomic Concept

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

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

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

Forest Habitat

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

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

Forest habitat from the trail.

It takes a Community

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

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

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

Botany Rocks

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

First rocky outcrop we saw on the trail.

Adapt 

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

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

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

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

Family Ties

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

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

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

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

Botanical Controversy 

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

Deadly Problem

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

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

Stand of mostly death camas with larkspur and paintbrush.

Microhabitat

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

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

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

Unusual section of Cone Peak bloom— rocky microhabitat.

Drawing Connections

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

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

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

More blooms on Cone Peak.

Ethnobotany 

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

The Summit

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

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

View from the summit.

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

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

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

Resources

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

Hike with a Fish Biologist

Looking up at the forest in the Valley of the Giants.

When hiking along a forest trail, fish are probably not the first thing that comes to mind.  What do fish have to do with forests? And forests with fish? 

It may not be obvious, but the connection between land and water is deep—as deep as a pool scoured by the fast action of water cascading over a fallen log. So when I met with Tony Spitzack, Fish Biologist for BLM, at the Valley of the Giants Outstanding Natural Area for a short hike through a late-successional forest, I was excited to discuss these connections.  

The Hike

  • Trailhead: Valley of the Giants Trailhead
  • Distance: 1.4 miles
  • Elevation Gain: about 500 feet
  • Details: Traveling to the Valley of the Giants trailhead requires patience. It is a long drive on logging roads. I recommend that you check with the district BLM office to check for closures and directions. However, the trailhead is well signed and there is a good amount of parking available.

Timber Time

The Valley of the Giants (VOG) Outstanding Natural Area lies just outside the deserted timber company town of Valsetz, west of Falls City in the Oregon Coast Range.

Small logging towns were once common in Oregon’s coastal mountains. As the logging industry of the late 19th and early 20th century boomed, and people were needed to support the growth. Then by the late 20th century, the industry changed, timber production declined, and the labor force required for logging was diminished. However, by this time, much of Oregon’s Coastal forests had already been logged.

As for Valsetz, In 1984, it was shut down and many of the buildings removed. It is now a tree farm.

Out with the Old

The impacts of this era of logging can still be seen today—the forests were altered and so were the rivers. Most of Oregon’s coastal old-growth was removed during that time. Coastal streams were channelized and disconnected from flood plains. Logs and natural barriers were removed and streams were cleared for transportation. According to Tony, splash dams were commonly used back then. A splash dam is a temporary wooden dam built to raise water levels enough to float timber downstream to sawmills. Few places remain in the Oregon Coast Range that didn’t experience these impacts.

The Valley of the Giants (VOG) is one of these few places. The VOG is a 51-acre piece of land that has never been logged. There is also no record of splash damming on the North Fork of the Siletz that runs through the VOG either (though it is unlikely that it has been untouched). Making the VOG a special place— a functional late-successional forest in a sea of mostly secondary growth. 

More primary growth trees standing tall in the Valley of the Giants.

Something Fishy

As we started along the trail, Tony told me a bit about the fish populations that inhabit the area. The trail crosses over the North Fork of the Siletz River, which supports chinook, summer steelhead, and cutthroat trout, among other fish species.  

Each species of fish found in the Siletz has a unique ecological niche. This means that they have different habits and take advantage of different habitats within the stream. For example, some species (or variants of species) of fish migrate to the ocean, like Chinook and Steelhead, while others will take up residence in freshwater streams or migrate to estuaries, like most cutthroat trout. 

According to Tony, the summer steelhead run is really unique. He said, “the North Fork of the Siletz River is the only native summer steelhead run within the Oregon Coastal steelhead population.” The reason being that downstream a boulder cascade waterfall, Siletz Falls, blocks the passage of steelhead during high winter flows. So the steelhead adapted to a summer run, creating a genetically unique population from winter steelhead. 

Cutthroat trout also distinguish themselves from other fish by their home range. One of Tony’s jobs is to determine upper fish limits for streams. Compared to other species, he finds cutthroat very high upstream. Making use of upper reaches of streams allows cutthroat to avoid competition with other fish and take advantage of this otherwise unreachable habitat. However, cutthroat populations sometimes move so high upstream that they become isolated when flows become low in these upper reaches.  Tony has even found cutthroat trout that have permanently lost connectivity with their home stream. What happens to these populations is still being studied.

Looking down at the North Fork of the Siletz River.

Riparian Feast

Gradually Tony and I made our way through the forest and closer to the stream edge. The corridor of vegetation along the stream is known as the Riparian area. Besides large conifer trees, deciduous trees and shrubs dominated the area.

Here, Tony reached out for a salmonberry to munch on. 

Besides providing delicious berries to eat, salmonberry and other deciduous trees and shrubs provide food to fish.  No, fish don’t eat the plants or berries, but leaves in various states of decay provide food for the insects and other invertebrates that feed fish. A large amount of energy is supplied by the riparian area through leaf fall to the aquatic food web in this way. 

An dark red salmonberry ready to be eaten.

Heterogeneity 

Tony also talked about how the riparian area also influences light availability and stream productivity.  He mentioned a study that showed that when riparian plants are removed, productivity increases, making more energy available to the stream food web. However, the result is only temporary.  In the years that follow, fast-growing riparian plants grow and shade out the stream reducing productivity significantly.  [1]

According to Tony, patchy cover that you get in a natural forest cycle is best. Tree fall creates patches of light to increase productivity, while other trees provide shade and other habitat needs to fish. A dynamic system is a resilient system, as change is a natural part of the ebb and flow of the ecosystem. 

Riparian trees and shrubs along the North Fork of the Siletz River.

Sort It Out

Shortly after crossing a bridge over the N. Fork of the Siletz, Tony and I reached the junction for the lollipop loop. We followed the trail to the right and soon reached a short spur that led to a small side channel.

Tony told me that he had scouted out the area earlier and decided the small channel was a great place to look at sediment sorting. The velocity of water in a stream determines how sediments sort. Smaller sediments, like sand, will move even in low velocity conditions, while gravel requires a higher velocity. Boulders are often found in the highest reaches of a stream system because they require higher velocity flows.  

Even in our small side channel, you could see sorting take place. Bigger rocks remained in the center, or thalweg, of the straight channel where the water traveled fastest. While smaller sediments collected along the edges where the water was moving slower. 

The way that sediments sort themselves within a channel of water is really important to fish. In particular, salmonid species rely on gravel beds for spawning habitat. Fish need complex streams with sediments that are sorted to create these gravel beds and other features, like pools, where sediments are scoured away. 

The small side-channel with sorted sediments.

Falling Logs

Tony placed a stick in the small channel to block the flow. He explained that as water poured over the stick it would start to wear away sediment creating a pool below it. He then moved the stick so that it entered the stream at 45 degrees. In this case, Tony said, “water hits the stick and turns perpendicular to the stick,” such that it will be directed toward the opposite bank and may undercut the bank.   

These same processes occur on a larger scale in big streams and rivers as well and are important to fish. Undercut banks provide excellent cover for fish, while pools are for resting and keeping cool. In addition, logjams trap sediments and aid in the formation of gravel beds needed for spawning. “The more you put wood in, the more dynamic the system is,” said Tony.  And the more dynamic the system, the greater the complexity and availability of different habitat features for fish. 

Log fall on the forest floor.

Networking

As we continued around the forested loop, I looked around at all the trees, shrubs, and forbs, and at the down logs with new growth sprouting from their rotting bodies— and I felt admiration.  These trees supplied so many benefits to fish—shelter, food, and rearing habitat—was there any reciprocation? 

Tony pointed at one of the larger down logs with shrubs and seedlings growing out it. He explained how the trees of the forest are connected by a mycorrhizal fungi network. Fungi gather nutrients and water from the soil and pass it to trees in exchange for carbon-rich sugars produced by trees through photosynthesis. 

This network reaches anywhere the forest grows, even into the nursery log before us. There are studies that show, shared Tony, that the mycorrhizal network is vast and far-reaching and fungi will carry nutrients and water over large distances in order to get the carbon they need. So even though we were a good distance from the river’s edge, our seedlings could benefit from the water and nutrients carried through the network— from the stream to the riparian area to our rotting log to our seedling. 

Shrubs and seedlings growing out of a nurse log.

Fish Feed Forests

As mentioned earlier, many of the fish that live in Oregon’s rivers migrate to the ocean to grow. Later, they return to their natal stream to spawn and die. Consequently, anadromous fish end up supplying marine nutrients to their freshwater home—a benefit to the ecosystem overall. 

This may seem like a small contribution, but when you consider the extent of Oregon’s native salmon runs—historically numbering in the 10s of millions— the magnitude of the transfer of nutrients is substantial. 

And don’t forget the mycorrhizae! The effects of these nutrients are far-reaching. Tony told me, “some people have suggested trees will grow three times as fast on a stream that has salmon coming back to it.” Clearly, fish do help trees.

Lost Fish

However, as we dammed and altered our river systems during the 20th century, salmonid populations plummeted. And the nutrient connection between marine and freshwater ecosystems was diminished greatly. 

Of course, that only counts fish that western culture historically paid any attention to. Tony told me about recent efforts to study Pacific Lamprey— a parasitic species with a tube-like body and round suctioning mouth-part—that has largely been ignored by western science. He has been using eDNA methodology to better understand the distribution of Lamprey in the Marys Peak Field Office. This non-invasive technique allows researchers to gather water samples from a body of water and run tests to see if the DNA of a fish is present in the water. 

We don’t know what Pacific Lamprey fish populations were historically. We don’t even know what they are now. But we do know that Pacific Lamprey were probably just as important, if not more important ecologically, as other anadromous fish.  Pacific Lamprey have high-fat content and are easier to prey on than salmonids. Tony suggested, “by biomass, they are probably more important.”

Either way, when considering the loss of salmonid and lamprey populations— “the amount of nutrients we have lost from the riparian area is astronomical,” said Tony. 

Disturbed 

As we finished the loop, crossed over the river and made our way back to our cars, I asked Tony what he thought were the biggest issues when it comes to fish. He was quick to respond: “The biggest thing humans have an issue with is that resilience depends on disturbance and not stability.”

I asked him to elaborate. He explained that humans want things to be neat and tidy—we don’t want ecosystems to change. We often see it as a bad thing. For example, landslides or forest fires are often seen as purely negative forces on an ecosystem. But the truth is, though perhaps destructive locally, the changes that result are overall positive. 

Tony mentioned a study that looked at landslide-prone areas. At first glance, you might think that landslides would be bad for fish. However, in this study, it was found that these areas not only had three important habitat types (spawning, summer -earing, and winter-refuge habitat) but connectivity between the habitat types. [2] Why? Because the landslides likely provide new materials, like sediment and wood to the stream, developing all the different habitat requirements Coho need, and in close proximity.

This is how nature works. Fish (and other wildlife) thrive in dynamic, heterogeneous environments. 

A bridge over the North Fork of the Siletz River.

Lessening Our Impact

Of course, not all disturbances are positive. Too much change can throw natural systems out of balance. So before we parted ways, I asked Tony what he thought about human disturbances in the lives of fish and forests. 

Tony offered a hopeful stance. He talked about how we have learned a lot about how to lessen our impacts on ecosystems over the years, while still benefiting from the products and services they provide. He used the example of roads.  We need roads, he explained, to access timber, recreation areas like the VOC, and other natural resources. But roads are going to impact the environment. However, not all roads are equal. Roads are now built with cross drains to divert runoff before reaching the stream, allowing sediment to settle in vegetated side-slopes. Also, new road construction now incorporates large culverts that allow streams to freely flow and fish to move up and downstream. Roads are necessary if we want access, according to Tony, but that doesn’t mean there isn’t a lot we can do to improve them and mitigate their impact.  

Driving Home

As I drove the many miles of gravel roads back to civilization, I thought a lot about what Tony said. 

It is apparent that humans are connected by a complex transportation network— we can easily see and experience this connection, as I did on the dusty, bumpy ride home. But there is another complex network that we are a part of that often remains hidden—the web of life. Like fish and trees in a forest—we too are dependent on the natural world—ecosystems provide us with clean air, water, food, medicine, and many more products and services.

So what do we do? We lay bare these connections. We study them and respect them. And by doing so, we build better roads—and perhaps a better world.

Tony Spitzack is a Fish Biologist with the Bureau of Land Management in the Northwest Oregon District.  Tony has also worked as a Natural Resource Technician for the Forest Service. He has a Masters Degree from Eastern New Mexico University and studied marine ecology at Washington State University Vancouver.

References

  1. Warren, D.R., Keeton, W.S., Bechtold, H.A. et al. Comparing streambed light availability and canopy cover in streams with old-growth versus early-mature riparian forests in western Oregon. Aquat Sci 75, 547–558 (2013). https://doi.org/10.1007/s00027-013-0299-2
  2. Beeson, Helen & Flitcroft, Rebecca & Fonstad, Mark & Roering, Josh. (2018). Deep‐Seated Landslides Drive Variability in Valley Width and Increase Connectivity of Salmon Habitat in the Oregon Coast Range. JAWRA Journal of the American Water Resources Association. 10.1111/1752-1688.12693.

Hike with a Terrestrial Wildlife Biologist

Looking onto Crabtree Lake.

The soft, spongy earth sinks and swells beneath my feet. Branches and needles tower overhead from trunks of various sizes and shapes, diffusing the light and casting shadows. The edges of grasses and herbs slip past my ankles, while shrubs tickle my things and hips. All the while an orchestra of whistles and sing-song sounds float on the wind, and a bouquet of sweet and musty smells rise and fall from the ground. Step, climb, dip, and try not to trip—this is what it is like to hike through a forest. 

When I met up with Corbin Murphy, BLM Wildlife Biologist, at the Crabtree Lake Trailhead, I knew that I was in for an adventure. The plan was to follow a trail down into the Crabtree Lake Valley, and then bushwack into the woods to reset some camera traps that needed tending to. We would eventually make it down to Crabtree Lake to one of the oldest forests in Oregon. I knew that walking would be a bit rough, but the payoff was worth it. I was right.

Corbin Murphy checks on his Beaver Dam Analog in the meadow.

The Hike

  • Trailhead: Crabtree Valley Trailhead
  • Distance: 4-5 miles
  • Elevation Gain: about 900 ft
  • Details: Roads to the trailhead are gravel but in decent condition. The last half mile of road is rough, but I made it with my Honda Civic. The usual route for this hike follows a decommissioned road down. Take a sharp right once you reach a road and follow it up to Crabtree Lake.

Diverse Species 

Entering a forest should be a rich, multisensory experience—an orchestra of sights, sounds, and scents.  It should be a tangled web of life! Complex ecosystems are not only more aesthetically pleasing, but they also tend to be resilient and functional. 

Paying attention to the diversity of species in an ecosystem is an important part of being a wildlife biologist. So, as Corbin and I began our hike along an old decommissioned road heading down toward Crabtree Valley, he was on high alert for the sights and sounds of the forest. It didn’t take long before we started talking about the different plants and animals we were seeing and hearing on the trail. 

Sounds of Life

Listening for birds was of particular interest to Corbin. He pointed out the high pitched electronic sounding whistle of a varied thrush and two-note chirp-chirp of a flycatcher. Because many birds are shy and difficult to spot in a forest, wildlife biologists often use bird calls to count birds instead of relying on visual identification. 

As part of his work, Corbin shared how he has been participating in breeding bird point-count surveys recently.  To conduct this kind of survey you drive along a transect an hour before dawn, stopping every half-mile for two minutes to listen, and identify bird calls. Point-counts are useful for biologists because they give us a better idea of what species are present in an ecosystem, and over time can see declines in specific species populations as well.

Green Stuff

In addition to birds, the variety of plant life also attracted our attention. Corbin pointed out several species of wildflowers, shrubs, and trees—you know, all the pretty green stuff.

It is easy to appreciate the importance of green stuff (a.k.a. plants) to an ecosystem.  From an early age, we learn that plants provide oxygen to breathe and food to eat. But not all plants are equal. Like animal species, each species of plant has its own role to fulfill in the ecosystem. In some cases, providing special benefits to select species. Thus, we need a diversity of plant life to support the diversity of life in an ecosystem. 

When it comes to conifer forests, less abundant deciduous trees and shrubs play a disproportionately large role in supporting the ecosystem. According to Corbin, conifer needles are generally not very nutritious. They have a low energy density, making them unable to support many invertebrate species. In contrast, deciduous trees and shrubs make a lot more energy available to support an abundance of species.

Deciduous trees along the trail.

Biological Desert

According to Corbin, a forest is more than just trees. A forest should have an understory of shrubs and forbs. In a natural system, stochastic disturbances, like forest fires, allow for the establishments of an understory.  High-density tree plantations do not. Corbin explained, “shrubs and forbs compete with seedlings. So they will establish, and they can dominate a site for anywhere from 30 to 300 years.” This stage of the forest is called “early seral” and is an important stage of forest development. 

However, in a tree plantation, this long period of competition is undesirable. Instead, a more profitable high-density forest is established, and the early seral stage of forest development is shortened or eliminated.  This creates “a biological desert,” said Corbin, “You have conifer trees and hardly any understory—any vegetation at all. You can literally count the number of plants and animals on one hand.”

That is why managing forests, like that surrounding Crabtree Lake, requires an eye for biodiversity. Forest density and early seral species should be considered. We don’t just need a bunch of any kind of plant, but we need an assortment of plants.

Look-Alike

Of course, even between deciduous understory trees, diversity of species is important. When hiking through a forest, it is easy to be blind to plant diversity. Everything can seem nondescript in a wash of greenery. But with a keen eye, even close look-alike species can be distinguished from one another.  

As we walked through a tunnel of deciduous trees and shrubs, Corbin pointed out a couple of look-alike pairs of species hidden in the foliage. 

One of the pairs that sat side-by-side was the Vine Maple and Rocky Mountain Maple also called Douglas Maple. Though very similar looking in size and general shape, vine maples tend to have more lobes, usually nine, than Rocky Mountain Maple, usually three.  Also, the Rocky Mountain Maple’s leaves have sharply toothed margins, while the Vine Maple’s leaves’ margins are doubly toothed.

Red Alder and Sitka Alder were another pair of look-alikes found on the trail. Again, though similar looking at first glance, the growth form of the Red Alder is straighter and taller, while the Sitka Alder is shorter and more shrubby.  Also, if you look closely at the leaves, the Red Alders’ leaf margins roll under slightly, while the Sitka Alders’ leaf margins are sharply toothed. 

All this to say, there are a lot of different kinds of green-stuff in a forest. 

Rocky Mountain Maple leaf overlaid with Vine Maple Leaf.

A Special Place

Before dropping down toward the lake, Corbin and I stopped to look down at where we were headed. Corbin explained that we were about to enter a really special place. Perhaps one of the oldest forests in Oregon, the Crabtree Lake Valley, and surrounding areas, are all part of the Crabtree Valley Complex—“An Area of Critical Environmental Concern (ACEC) due to its outstanding geological, recreational, and ecological value.”

Crabtree Valley was created during the last ice age. Glaciers carved out large amphitheater-like valleys, called cirques, which protected much of the forest from fire for perhaps 1,000 years. Later, for whatever reason, it remained unlogged.  Making it a perfect example of a late-successional forest and refuge for species, like the Northern Spotted Owl. 

So when the BLM acquired the land in the 1980s, it fell under ACEC status and a management plan was put in place in order to protect its values. Which brings us to today where it is still under a resources management plan as a late-successional reserve. 

View into the meadow with protective rock.

Management to Protect

One of the ways the BLM has been working to meet the goals of the resource management plan is by reducing roadways in the area. Though some areas within the Crabtree Lake Complex were never logged, logging was still rampant in the region. In fact, the first part of our hike was on an old logging road through an area that was probably logged in the 70s or 80s.

So in order to enhance and restore what we might expect from a late-successional reserve, the BLM decommissioned most of the roads, ripped them up, put in waterbars, and took out culverts—all efforts to restore the natural functions of the forest. 

Give a Hoot

Eventually, we made our way down to the lower valley floor and into the late-successional forest reserve. Here we took a sharp left onto another road Corbin said he usually uses to access the property. He also told me that the road is where the BLM does surveys for Northern Spotted Owl. Every half-mile along the road is a survey station where a biologist will stop for 10 minutes to call and listen for spotted owls. 

There are two pairs of spotted owls reported within the watershed, Corbin said, because “the habitat is so great in this area.” This is unusual because spotted owls usually need a 1.2 mile home range in the Cascades, but these nesting pairs are only about a half-mile apart. Not only that, but last year the pairs each had two juveniles. Which is remarkable because, as Corbin explained, “other than that, there was zero reproduction in spotted owls from Sweet Home in the BLM up to the Columbia River.” 

Wear Layers 

Continuing down the road, the dynamics of the forest opened up— there were tall douglas-fir trees and hemlock; open areas with shrubs and smaller trees; and snags and down logs. 

 “One of the big things about late-successional forests too is the structure,” said Corbin.  You want to see “horizontal and vertical heterogeneity” in a late-successional forest.

Basically, a forest like the one we were observing, starts with a lot of Douglas-fir, but then over the next hundred years, holes open up in the canopy that allows shrubs and shade-tolerant trees, like hemlock, to grow and fill in gaps.  

This development of structure is important because it creates habitat for wildlife. A forest that lacks diverse forest structure is simply not conducive to the wildlife that needs late-successional forest.

Corbin told me about a transect study that looked at how flying squirrels fared when there were big trees, but no holes for shrubs and smaller trees available for the development of an understory. The squirrels had the big trees they needed for food and nesting, but there was not enough cover for them to avoid predation. Needless to say, the outcome wasn’t great for the squirrels 

Highly structured forest observed along the road.

A Rotten Heart

At one point, Corbin and I came across a down tree with heart rot. Which brings me to another component of late-successional forest that adds to its complexity— dead stuff.

If the down tree with heart rot was actually standing, or a snag, it would provide habitat for cavity nesters like woodpeckers. As a large down log, it creates habitat for hundreds of invertebrates, bacteria, and fungi, as well as amphibians. 

The importance of dead trees cannot be overemphasized. In fact, often land managers create snags by girdling trees in an attempt to mimic the natural process of snag formation. Unfortunately, according to Corbin, it generally doesn’t work very well.  The natural process is slow, possibly taking a couple of hundred years for a snag to form. There really isn’t a quick way to recreate that. 

In addition, Oregon slender salamanders, a species of concern, rely on the late-successional forest for large down wood. This species is endemic to Oregon and is doing O.K. right now, but as timber harvesting continues to produce young 20-30 years old forests, things could get dicey. Less large down wood means less of an important microhabitat that Oregon slender salamanders need to survive.

This is why on federal public lands, Corbin explained, “we are trying to institute measures to have leave trees, and these are the legacy trees from the previous cohort, and those are the ones that have all the lichens and bryophytes—create a little refugium—and those eventually become snags and fall over.” 

Downed Log with heart rot as seen on the trail.

Leave it to Beaver

Not long after passing the downed log, Corbin and I headed off-trail to check on a beaver dam analog (BDA) that was put in last fall.  As we climbed through the underbrush, Corbin explained that beavers were historically present in the wet meadow we were about to visit, pooling the water and creating a much larger lake. We even some old beaver sign to confirm it.

However, when roads were constructed in the area, the beavers disappeared. Corbin hypothesized that they could have been trapped. Since then, trees have started to encroach into the wet meadow, altering the historically flooded area and shrinking the lake.  

Then, a couple of years ago, Oregon Department of Fish and Wildlife and BLM joined forces in an effort to reintroduce beavers into the area.  Several beavers were released into the watershed. But they didn’t stay. 

Now, the BLM is working on a soft release program in the hopes that the next group of beavers they introduce won’t go away.  That is why the BLM constructed the BDA—in an effort to make the meadow homier. Once established, beavers, a keystone species, will naturally alter the ecosystem; hopefully, restoring the meadow to historical conditions.

A Beaver Dam Analog (a.k.a—fake beaver dam) in the wet meadow.

Fishing for Fishers

After visiting the BDA, Corbin and I continued a bit further down the road before making our way back into the woods again. This time we bushwacked our way to one of the camera-traps Corbin needed to reset. The camera traps were set up as part of a Forest Carnivore Research Project started by Katie Moriarty from Oregon State University. The BLM adopted a project grid area, and are working on tracking the carnivores that visit each camera trap site. 

The overall goal of the projects is to determine if Pacific Fishers are present in the Western Cascades. Historically, Corbin shared, Pacific Fishers ranged from California up through British Columbia. But their range has shrunk in Oregon over the years and now there is no record of Pacific Fishers anywhere north of Eugene. Later, as part of the carnivore project, Fishers will hopefully be reintroduced into areas like the Crabtree Lake Valley. 

As Corbin worked to reset the camera trap and bait it, I asked him about why the reintroduction of Fishers is important. He explained that Fishers are candidate species for ESA listing, which makes them important in the eyes of the government.  Candidate species are in danger of extinction in at least part of their range.

Species extinction is a concern because, as mentioned earlier, each species has a role in the ecosystem. Pacific Fishers are top predators. They help regulate populations of organisms that sit below them in the food chain. They are also opportunistic feeders and primarily prey on small mammals, including squirrels and even porcupines. Thus the loss of Fishers could have ripple effects on the forest food web—allowing porcupine populations to increase, for example, which could lead to excessive damage to trees they feed on.

Carnivore Project bate opposite camera trap.

Forest Walking

After resetting the first camera trap, we did some serious bushwhacking up to the next one before heading down to crabtree lake.  As we made our way to the lake, I was taken aback by the grandeur of the forest. I felt small beside the mammoth-sized trees, but at the same time, perfectly natural walking across a huge moss-covered log. We were really in the thick of the forest.  

Here we did the forest dance—climbing, ducking, and trying not to trip. We saw more life, including a small salamander hiding amongst a pile of old deadwood. We talked about huckleberries that would ripen in late summer. And craned our necks looking up at the tallest trees in the forest. 

During the last leg of our hike, the biodiversity of habitat and species was all around us—the promise of spotted owls, flying squirrels, and future fisher. This is what hiking in a forest is all about! 

Corbin doing the “forest dance” as we bushwhacked our way to Crabtree Lake.

Corbin Murphy is a Wildlife Biologist for the Salem District of Bureau of Land Management. He has been with the BLM for 11 years and currently works in the Cascades Field Office. He has also worked for the U.S. Forest Service.