What Really Happened at the Oroville Dam Spillway?

Video Statistics and Information

Video
Captions Word Cloud
Reddit Comments

This was actually a very good video, as all of the practical engineering videos are

👍︎︎ 32 👤︎︎ u/_stonks_only_go_up_ 📅︎︎ May 30 2021 🗫︎ replies

I live in area, south of the dam, that was evacuated just in case everything went wrong. So hearing about this is awesome.

👍︎︎ 18 👤︎︎ u/Father_fist_cuervo 📅︎︎ May 30 2021 🗫︎ replies

Great video, would like to have seen/learned more about the repair

👍︎︎ 11 👤︎︎ u/postdochell 📅︎︎ May 30 2021 🗫︎ replies

awesome video. this is the kinda stuff that really makes youtube shine.

20 years ago we wouldn't have had access to this kinda informative content.

👍︎︎ 10 👤︎︎ u/turbodude69 📅︎︎ May 30 2021 🗫︎ replies
Captions
In February 2017, concrete slabs in the  spillway at Oroville Dam failed during   releases from the floodgates, starting a chain  of events that prompted the evacuation of nearly   200,000 people downstream. The dam didn’t  fail, but it came too close for comfort,   especially for the tallest structure  of its kind in the United States.   Oroville Dam falls under the purview of  the Federal Energy Regulatory Commission,   in a state with a progressive dam safety program  and regular inspections and evaluations by the   most competent engineers in the industry. So how  could a failure mode like this slip through the   cracks, both figuratively and literally? Luckily,  an independent forensic team got deep in the weeds   and prepared a 600 page report to try and find  out. This is a summary of that. I’m Grady and this   is Practical Engineering. In today’s episode,  we’re talking about the Oroville Dam Crisis. Oroville Dam, located in northern California,  is the tallest dam in the United States   at 770 feet or 235 meters high. Completed in  1968, and owned and operated by the California   Department of Water Resources, every part  of Oroville Dam is massive. The facility   consists of an earthen embankment which forms  the dam itself, a hydropower generation plant   that can be reversed to create pumped storage,  a service spillway with 8 radial floodgates,   and an emergency overflow spillway. The  reservoir created by the dam, Lake Oroville,   is also immense - the second biggest in the state.  It’s part of the California State Water Project,   one of the largest water storage and delivery  systems in the U.S. that supplies water to more   than 20 million people and hundreds of thousands  of acres of irrigated farmland. The reservoir is   also used to generate electricity with over  800 megawatts of capacity. Finally, the dam   also keeps a reserve volume empty during the wet  season. In case of major flooding upstream, it can   store floodwaters and release them gradually over  time, reducing the potential damage downstream. No dam is built to hold all the water that could  ever flow into the reservoir at once. And yet,   having water overtop an unprotected embankment  will almost certainly cause a breach and failure.   So, all dams need spillways to safely release  excess inflows and maintain the level of the   reservoir once it’s full. Spillways are often the  most complex and expensive components of a dam,   and that is definitely true at Oroville. The  service spillway has a chute that is 180 feet or   55 meters wide and 3,000 feet long. That’s nearly  a kilometer for the metric folks. Radial gates   control how much water is released and massive  concrete blocks at the bottom of the chute, called   dentates, disperse the flow to reduce erosion as  it crashes into the Feather River. This spillway   is capable of releasing nearly 300,000 cubic feet  or 8,000 cubic meters of water per second. That’s   roughly an olympic-sized swimming pool every  other second, which I know is not that helpful   in conceptualizing this incredible volume. If  you somehow put that much flow through a standard   garden hose, it would travel at 15% of the speed  of light, reaching the moon in about 9 seconds.   How’s that for a flow rate equivalency? But even  that is not enough to protect the embankment. Large dams have to be able to withstand  extraordinary flooding. In most cases,   their design is based on a synthetic (or made  up) storm called the Probable Maximum Flood,   which is essentially an approximation of the most  rain that could ever physically fall out of the   sky. It usually doesn’t make sense to design  the primary spillway to handle this event,   since such a magnitude of flooding is unlikely to  ever happen during the lifetime of the structure.   Instead, many dams have a second spillway,  much simpler in design - and thus less   expensive to construct - to increase their  ability to discharge huge volumes of water   during rare but extreme events. At Oroville, the  emergency spillway consists of a concrete weir   set one foot above the maximum operating  level. If the reservoir gets too high and   the service spillway can’t release water  fast enough, this structure overflows,   preventing the reservoir from reaching  and overtopping the crest of the dam. Early 2017 was one of northern California’s  wettest winters in history with several major   flood events across the state. One of those storms  happened in February upstream of Oroville Dam.   As the reservoir filled, it became clear to  operators that the spillway gates would need   to be opened to release excess inflows.  On February 7, early during the releases,   they noticed an unusual flow pattern  about halfway down the chute. The   issue was worrying enough that they decided to  close the gates and pause the flood releases   in order to get a better look. What they saw  when the water stopped was harrowing. Several   large concrete slabs were completely missing  and a gigantic hole had eroded below the chute. There was a lot more inflow to the reservoir in  the forecast, so the operators knew they didn’t   have much time to keep the gates closed while  they inspected the damage, and no chance to try   and make repairs. They knew they would have  to keep operating the crippled spillway. So,   they started opening gates incrementally to test  how quickly the erosion would progress. Meanwhile,   more rain was falling upstream, contributing to  inflows and raising the level of the reservoir   faster and faster. It wasn’t long before the  operators were faced with an extremely difficult   decision: open more gates on the service spillway  which would further damage the structure or let   the reservoir rise above the untested emergency  spillway and cascade down the adjacent hillside. Several issues made this decision  even more complicated. On one hand,   the service spillway was in bad shape, and there  was the possibility of the erosion progressing   upstream toward the headworks which could result  in an uncontrolled release of the reservoir.   Also, debris from the damaged spillway  was piling up in the Feather River,   raising its level and threatening to  flood out the power plant. Finally,   electrical transmission lines connecting the  power plant to the grid were being threatened   by the erosion along the service spillway. Losing  these lines or flooding the hydropower facility   would hamstring the dam’s only backup for making  releases from the reservoir. Operators knew that   repairing the spillway would be nearly impossible  until the power plant could be restored. These   factors pointed towards closing the spillway  gates and allowing the reservoir to rise. On the other hand, the emergency spillway  had never been tested, and operators weren’t   confident that it could safely release so much  water, especially after witnessing how quickly   and aggressively the erosion happened on the  service spillway nearby. Also, its use would   almost certainly strip at least the top layer  of soil and vegetation from the entire hillside,   threatening adjacent electrical transmission  towers. A huge contingent of engineers   and operations personnel were all hands on  deck, running analyses, forecasting weather,   reviewing geologic records and original design  reports trying to decide the best course of   action. Of course, this is all happening  over the course of only a couple of days   with conditions constantly changing and no one  having slept, further complicating the decision   making process. Operators worked to find a sweet  spot in managing these risks, limiting releases   from the service spillway as much as possible  while still trying to keep the reservoir from   overtopping the emergency spillway. But, every new  forecast just showed more rain and more inflows. Eventually it became clear to operators that  they would have to pick a lesser evil: Increase   discharges and flood the powerhouse or let the  reservoir rise above the emergency spillway.   They decided to let the reservoir  come up. The morning of February 11,   about four days after the damage was  initially noticed, Lake Oroville rose   above the crest of the emergency spillway  for the first time in the facility’s history.   Almost immediately, it was clear that  things were not going to go smoothly. As it flowed across and down the natural hillside,  water from the emergency spillway began to   channelize and concentrate. This quickly  accelerated erosion of the soil and rock,   creating features called headcuts, which are a  sign of unstable and incising waterways. Headcuts   are vertical drops in the topography eroded by  flowing water, and they always move upstream   oftentimes aggressively. In this case, upstream  meant toward the emergency spillway structure,   threatening its stability. This hillside  was a zone many had assumed to be solid,   competent bedrock. It only took a modest flow  through the emergency spillway to reveal the   true geologic conditions: the hillside was  composed almost entirely of highly erodible soil   and weathered rock. If the headcuts were  to reach the concrete structure upstream,   it would almost certainly fail, releasing  a wall of water from Oroville Lake that   would devastate downstream communities.  Authorities knew they had to act quickly. On February 12, only about a day and half  after flow over the emergency spillway began,   an evacuation order was issued for downstream  residents, displacing nearly 200,000 people   to higher ground. At the same time, operators  elected to open the service spillway gates to   double the flow rate and accelerate the lowering  of the reservoir. The level dropped below the   emergency spillway crest that night, stopping the  flow and easing fears about an imminent failure.   Two days later, on Valentine’s Day, the evacuation  order was changed to a warning, allowing people to   return to their homes. But there was still more  rain in the forecast, and the emergency spillway   was in poor condition to handle additional  flow if the reservoir were to rise again.   California DWR continued discharging through the  crippled service spillway to lower the reservoir   by 50 feet or 15 meters in order to create  enough storage that the spillway could be taken   out of service for evaluation and repairs.  The gates stayed open until February 27th,   nearly three weeks after the whole mess started,  revealing the havoc to the dam’s right abutment.   Water that started its journey as tiny  drops of rain in a heavy storm - funneled   and concentrated by the earth’s topography and  turbulently released through massive human-made   structures - had carved harrowing scars  through the hillside. But, how did it happen? Like all major catastrophes, there were a  host of problems and issues that coincided   to cause the failure of the concrete chute.  One of the most fundamental issues was   geologic. Although it was  well-understood that some   areas of the spillway’s foundation were not good  stuff (in other words, weathered rock and soil),   the spillway was designed and maintained as if  the entire structure was sitting on hard bedrock. That mischaracterization had profound  consequences that I’ll discuss.   As for how the spillway damage started, the  issue was uplift forces. How do concrete   structures stay put? Mostly by being heavy. Their  weight pins them to the ground so they can resist   other forces that may cause them to move.  But, water complicates the issue. You might   think that adding water to the top  of a slab just adds to the weight,   making things more stable. And that would be true  without cracks and joints. The problem with the   Oroville Dam service spillway chute was that it  had lots of cracks and joints, for reasons I’ll   discuss in a moment. These cracks allowed water to  get underneath the slabs, essentially submerging   the concrete on all sides. Here’s the issue  with that: structures weigh less under water,   or more accurately, their weight is counteracted  by the buoyant force of the water they displace.   So, being underwater already starts to destabilize  them, because it adds an uplift force. But,   concrete still sinks underwater, right? The  net force is still down, holding the structure   in place. That’s true in static conditions,  but when the water is moving, things change. We talk about Bernoulli’s principle a lot on this  channel, and he’s got something to say about the   flow of water in a spillway. In this case, the  issue was what happens to a fast-moving fluid when   it suddenly stops. Cracks and joints in a concrete  spillway have an effect on the flow inside. Any   protrusion into the stream redirects the flow. If  a joint or crack is offset, that redirection can   happen underneath the slab. When this happens, all  the kinetic energy of the fluid is converted into   potential energy, in other words, pressure. When  it’s 100% of the kinetic energy being converted,   we call it the stagnation pressure. See how the  level rises in this tube when I direct it into   the flowing water. The equation for stagnation  pressure is a function of velocity squared.   So, if I double the speed of flow in my flume,  I get four times the resulting pressure and thus   four times the height the water rises in my tube.  And the water in the Oroville spillway is moving   a lot faster than this. When this stagnation  pressure acts on the bottom of a concrete slab,   it creates an additional uplift force. If all  the uplift forces exceed the weight of the slab,   it’s going to move. That’s exactly what  happened at Oroville. And once one slab goes,   it’s just a chain reaction. More of the  foundation is exposed to the fast moving water,   and more of that water can inject itself  below the slabs, causing a runaway failure. Of course, we try to design around this problem.  The service spillway had drains consisting of   perforated pipes to relieve the pressure of  water flowing beneath the slabs. Unfortunately,   the design of these drains was a major reason for  the cracking chute. Instead of trenching them into   the foundation below the slabs, they reduced  the thickness of the concrete to make room   for the drains. The crack pattern on the chute  essentially matched the layout of the drains   beneath perfectly. So, in this case the drains  inadvertently let more water below the slab than   they let out from underneath it. The chute also  included anchors, steel rods tying the concrete to   the foundation material below. Unfortunately those  anchors were designed for strong rock and their   design wasn’t modified when the actual foundation  conditions were revealed during construction. The root cause wasn’t just a bad design, though.  There are plenty of human factors that played   into the lack of recognition and failure to  address the inherent weaknesses in the structure.   Large dams are regularly inspected, and their  designs periodically compared to the state of   current practice in dam engineering. Put  simply, we’ve built bigger structures on   worse foundations than this. Modern spillway  designs have lots of features that help to avoid   what happened at Oroville. Multiple layers  of reinforcement keep cracks from getting too   wide. Flexible waterstops are embedded into joints  to keep water from migrating below the concrete.   Joints are also keyed so individual slabs  can’t separate from one another easily.   Lateral cutoffs help resist sliding and keep  water from migrating beneath one slab to another.   Anchors add uplift resistance by holding  the slabs down against their foundation.   Even the surface of the joints is offset  to avoid the possibility of a protrusion   into the high velocity flow. All these are things  that the Oroville Spillway either didn’t have or   weren’t done properly. Periodic reviews of the  structure’s design, required by regulators,   should have recognized the deterioration  and inherent weaknesses and addressed   them before they could turn into such  a consequential chain of tribulations. As for the emergency spillway, the  fundamental cause of the problem was similar:   a mischaracterization of the foundation material  during and after design. Emergency spillways are   just that: intended for use only during a rare  event where it’s ok to sustain some damage.   But, it’s never acceptable for the structure  to fail, or even come close enough to failing   that the residents downstream have to be  evacuated. That means engineers have to be   able to make conservative estimates of how much  erosion will occur when an emergency spillway   engages. Predicting the amount and extent of  erosion caused by flowing water is a notoriously   difficult problem in civil engineering. It takes  sophisticated analysis in the best of times,   and even then, the uncertainty is still  significant. It is practically impossible to   do under the severe pressure of an emergency. The  operators of the dam chose to allow the reservoir   to rise above the crest of the emergency  spillway rather than increase discharges   through the debilitated service spillway,  trusting the original designer that it could   withstand the flows. It’s a decision I think  most people (in hindsight) would not have made. The powerhouse was further from flooding and  the transmission lines further from failing than   initially thought, and they eventually ramped  up discharges from the service spillway anyway,   after realizing the magnitude of the  erosion happening at the emergency spillway.   But, it’s difficult to pass blame too strongly.  The operators making decisions during the heat   of the emergency did not have the benefit  of hindsight. They were stuck with the many   small but consequential decisions made over  a very long period of time that eventually   led to the initial failure, not to mention  the limitations of professional engineering   practice’s ability to shine a light down  multiple paths and choose the perfect one. The forensic team’s report outlines many lessons  to be learned from the event by the owner of the   dam and the engineering community at large,  and it’s worth a read if you’re interested in   more detail. But, I think the most important  lesson is about professional responsibility.   The people downstream of Oroville Dam,  and indeed any large dam across the world,   probably chose their home or workplace without  considering too carefully the consequences of   a failure and breach. We rarely have the luxury to  make decisions with such esoteric priorities. That   means, whether they realized it or not, they  put their trust in the engineers, operators,   and regulators in charge of that dam to keep them  safe and sound against disaster. In this case,   that trust was broken. It’s a good reminder  to anyone whose work can affect public safety.   The repairs and rebuilding of the spillways at  Oroville Dam are a whole other fascinating story.   Maybe I’ll cover that in a future video. Thank  you for watching, and let me know what you think!
Info
Channel: Practical Engineering
Views: 3,164,151
Rating: undefined out of 5
Keywords: oroville dam, dam failure, spillway, floodgate, dentate, feather river, probable maximum flood, weir, reservoir, erosion, headcuts, stagnation pressure, Practical Engineering, Civil Engineering, Engineer, Grady
Id: jxNM4DGBRMU
Channel Id: undefined
Length: 18min 27sec (1107 seconds)
Published: Tue May 18 2021
Related Videos
Note
Please note that this website is currently a work in progress! Lots of interesting data and statistics to come.