Aviation Weather Safety Archives - Plane & Pilot Magazine https://cms.planeandpilotmag.com/article/proficiency/weather-flying/ The Excitement of Personal Aviation & Private Ownership Mon, 27 Sep 2021 13:20:50 +0000 en-US hourly 1 https://wordpress.org/?v=6.4.4 We All Need To Be Weather Geeks Now https://www.planeandpilotmag.com/article/need-weather-geeks-now/ Fri, 26 Jan 2018 17:13:40 +0000 http://www.planeandpilotmag.com/?post_type=article&p=29730 From airfactsjournal.com. Disruption is a hot topic these days, but the 16th century had plenty of it, too. In the span of a generation, the printing press (and the Protestant...

The post We All Need To Be Weather Geeks Now appeared first on Plane & Pilot Magazine.

]]>
From airfactsjournal.com.

Disruption is a hot topic these days, but the 16th century had plenty of it, too. In the span of a generation, the printing press (and the Protestant Reformation) fundamentally changed the way many Christians practiced their religion. Instead of listening to a priest interpret the Bible, the average churchgoer could read it in his native language and make up his own mind. This movement wasn’t universal, and it wasn’t without controversy, but for many, power had clearly shifted from the pulpit to the pews by 1585.

Aviation is experiencing a similar shift, although it probably won’t change the world in such a momentous way. Thanks to the ubiquity of internet access, the power of software, and the diversity of mobile devices, pilots no longer have to depend on a Flight Service briefer or TV meteorologist to explain the weather. We can now consult our iPads, which are overflowing with sophisticated weather products, from radar imagery and icing graphics to MOS forecasts and Skew-t diagrams. Weather forecasting has been democratized.

Most of us probably consider this change to be liberating, but it’s not without drawbacks. While apps like ForeFlight and Garmin Pilot can simplify the flight planning process, if we’re not careful they can also make it confusing. We are all our own Flight Service Stations now, forced to assemble weather information, evaluate it, and make a plan. Which sources can be trusted? What do they all mean? How much weather information is enough?

To answer questions like these, pilots need more than just a passing acquaintance with aviation weather. The differences between base and composite radar reflectivity, or the limitations of a cloud top forecast, are not just trivia. They determine whether our new tools are safety enhancements or distractions. Use them correctly, and it has never been easier to make a safe decision; use them poorly, and you can stumble into a deadly trap.

Unfortunately, a growing number of pilots don’t seem to be interested in such details. So at exactly the moment when we need to be weather geeks more than ever, pilots are tuning out. Why?

True weather geeks pay attention to the view outside as well as internet weather data

Teaching Weather Flying

Our current system of aviation education, both the curriculum and the instructors, deserves some of the blame. As the old saying goes, “if you measure it, you aspire to it,” and right now ground reference maneuvers get measured a lot more than weather decision-making skills. A pilot is much more likely to bust a checkride because their S-turns weren’t within ACS standards than because they read the upper air analysis incorrectly. Given this bias toward easily scored maneuvers, it’s no surprise they get the most attention during primary training lessons.

Learning about weather flying is also a lot more time-consuming (and thus more expensive) than most other subjects. A motivated student pilot can do eight landings in a one-hour lesson, but it’s hard to cram a lot of practical weather experience in the same amount of time. That happens beyond the practice area and often over dozens of lessons, as weather charts are compared to actual conditions aloft. With flight lessons costing over $200/hour at some schools, instructors are understandably focused on efficiency, so these finer points are often skipped over entirely.

Some instructors do cover weather in more detail, but even this sometimes devolves into scaremongering instead of thoughtful instruction. There might be stern lectures about the delay in datalink radar (true but overrated) or the requirement to call Flight Service for an “official briefing” (completely false—there’s no such thing). This is often blamed on old timers being curmudgeonly, but more likely it’s due to the instructor’s lack of experience with new technology. The skills required to be a great primary flight instructor are not the same as those required to be a great weather flying pilot, and the rapid pace of technological change means some instructors are simply not comfortable with the latest tools.

weather radar
We are all our own Flight Service Stations now, in the position to assemble weather information, evaluate it, and make a plan.

In my experience, the best weather instruction seems to occur during instrument training. This makes sense on a very basic level—instrument flying is all about managing the risks of flying near weather, so it’s the perfect time to geek out—but reserving serious weather training for this step is a major mistake. Most obviously, such a strategy does nothing to help the 50% of pilots who don’t have an instrument rating or to address the persistence of VFR-into-IMC accidents.

Just as importantly, pilots who ignore weather during private pilot training often end up less confident about cross-country flying in general and less willing to use their new pilot privileges for transportation. That’s a shame, because those longer trips are some of the best experiences you can have in an airplane.

Technology And Its Promises

All this talk about instructors should not obscure the role technology plays in changing our expectations and habits. Consider Google and its slow evolution from database to interactive assistant. Five or ten years ago, typing “what’s the weather?” would have produced a list of websites. From there, a searcher could compare various forecasts and charts, then come up with an answer. It was a process, and it took time and deliberation.

Type that same query into Google today and you’ll instantly be presented with an answer: local temperature, cloud cover and forecast. No need to click through multiple websites and no need to even enter your location. The search engine knows where you are and has decided for you what the best forecast is.

It’s a subtle change, but it has significant implications. Applying this simplistic approach to complex personal decisions is asking for trouble, but some pilots are trying. The clearest example of this trend is the current crop of Flight Risk Assessment Tool (FRAT) apps. These present a series of questions, from pilot experience to destination weather conditions, then score the flight based on the potential risk, complete with a red light/green light answer. It’s a fine concept, but it’s awfully messy in practice. The go/no-go decision, much like bond ratings in 2008, is not easily summed up by an algorithm.

A recent VFR flight reminded me how fuzzy these decisions are in the real world. I was flying home from Oshkosh in a Robinson R44, and the weather was interesting enough to create some uncertainty for this 500-foot AGL flight. The departure airport had clear skies and good visibility—a FRAT would have been flashing green—but conditions 150 miles south were more complicated, with showers, thunderstorms and 1,000-foot ceilings across Illinois and Indiana. Canceling would have been easy, but with a little knowledge and some studying, things looked less threatening. The storms were scattered, not stacked up in a line, and they weren’t building as the day wore on.

ForeFlight weather display
Managing missions when the weather is convective and fast-changing requires more than a knowledge of color coding. Where are storms likely to form next, where will you be in your flight when that happens, and what are your options then, are all critical questions best answered during preflight briefings but with an eye to making changes on the fly, if need be.

My decision to take off was not the result of an app or a simple radar picture; I departed because I knew enough about the overall weather system to suspect it would weaken as time went on. I’m not sure an app could have reached this decision on its own. An awareness of the big picture, combined with a willingness to be flexible, changed the classic go/no-go decision to “go a little bit and reevaluate 100 miles south.”

New technology definitely made the flight easier. My iPhone kept me updated on the latest weather conditions before takeoff, even while strolling the grounds at Oshkosh. After takeoff, the datalink weather on my iPad was extremely helpful for keeping the big picture in mind. Its most important role was to prevent me from being backed into a corner as conditions changed. As good as ForeFlight was, though, the best weather tool was the front window, and almost all of the flying involved nothing more than dodging ugly clouds. In the end, there was hardly a bump or a drop of rain, although I certainly didn’t fly direct.

The problem is, it’s simply impossible to simulate this kind of decision-making during a ground lesson or on a knowledge test. Neither textbooks nor technology can make flying safe or easy. That only comes with practical training, experience and discipline.

The word progress suggests improvement, but that doesn’t mean it’s always easy or straightforward. Aviation apps and new forecast tools are indeed progress, but we shouldn’t cede our authority over weather decision making just because the graphics are great. Neither should we use technology as an excuse to stop learning or become passive consumers of weather data. After all, we are drowning in data, not answers. Those must come from the pilot in command.


John Zimmerman is vice president of Sporty’s Pilot Shop’s Catalog Division. He regularly flies a Citabria, a Pilatus PC-12 and a Robinson R44 helicopter. You can find John’s blog on Air Facts at airfactsjournal.com.


Want more on pilot skills—from flight training to weather flying? Visit our Proficiency archives.

The post We All Need To Be Weather Geeks Now appeared first on Plane & Pilot Magazine.

]]>
To Go, Or Not To Go? https://www.planeandpilotmag.com/article/go-not-go-2/ Tue, 21 Feb 2017 10:49:20 +0000 http://www.planeandpilotmag.com/?post_type=article&p=25922 That's the (wrong) question.

The post To Go, Or Not To Go? appeared first on Plane & Pilot Magazine.

]]>
From airfactsjournal.com.

At its most basic, flying an airplane is a never-ending series of decisions. Is the airplane airworthy? What’s the weather like? Where’s that other airplane going? When should I turn base? Failing to ask these questions and make timely decisions is a serious mistake—one that will earn you a place in an NTSB report if you’re unlucky.

But there’s another decision-making error lurking out there, one that’s simultaneously more common and rarely discussed: We falsely view most aviation decisions as binary. The language of decision-making subtly reinforces this, with exhortations to “keep it simple” or “be confident.” What we end up with is a hopelessly unrealistic set of answers: yes or no, black or white.

We should know better. Flying is all about subtle clues, 50/50 decisions and shades of gray.

The most important example of this fallacy is the vaunted “go/no-go” decision, a topic that fills textbooks and flight lessons. A pilot looks at his airplane, his skills and the weather conditions, then decides whether it’s safe to fly the planned trip. Most of the time, this is presented as a simple yes or no question. (Heck, we’re as guilty as anyone here at Air Facts, with our interactive weather decision-making series of the same name.)

But making the go/no-go decision is hardly that simple. If it’s carefully considered, it’s really a series of questions with a variety of possible answers, more an essay than a true-false question. Here are four ways to expand your concept of the famous decision point.

Go Now Or Go Later/Earlier?

A great example of this more nuanced philosophy is the estimated time of departure (ETD). Viewed as a binary choice, that line of storms bearing down on the airport looks like an easy no-go. But what about leaving an hour earlier, before the storms get close, or leaving the next morning, when skies will be clear? That type of schedule flexibility is exactly what makes general aviation so fun and useful. While most pilots know this, many don’t take advantage of it. Don’t get so locked into your original plan that you fail to see attractive alternatives.

Go Direct, Or Go Around?

My multi-engine flight instructor used to occasionally challenge me during preflight planning by saying: “You have to go, show me a route that is safe.” It was an exaggeration—we never have to go—but he was an airline pilot by day and wanted me to get a feel for his decision-making mindset. While we may not have a transport category jet at our disposal, it’s amazing how often a flight can be completed safely and comfortably if you’re willing to deviate. On a 400-mile trip, even a 100-mile detour will be quickly forgotten if you have a smooth ride and get to your destination. Especially in the United States, we have incredible freedom to make up our own route and getting there is half the fun anyway.

Maybe you can’t go direct, but you can still go.

Go All The Way Or Go Part Of The Way And Stop?

I believe the age-old approach of “taking a look” gets a bad reputation. Certainly, naively blasting off and hoping low IFR conditions will magically disappear is a bad idea. But flying as far as the weather allows and then diverting can be smart and effective. Some days the only option is to fly up to a line of rain, land, and wait for it to pass. That might mean an hour or a day on the ground, but if it’s done with firm limits in mind and plenty of backup options, there’s no reason this can’t be another useful tool in the savvy pilot’s bag. Similarly, you can choose to go some of the way and turn around if it’s worse than forecast. It takes discipline, but Richard Collins has written before about how practical this approach can be.

Go Solo Or Take Another Pilot?

Here’s one that’s especially neglected by newer pilots. Some days, especially when the forecast involves thin cloud layers or gusty winds, the safe answer is no-go. But if you’re willing to push yourself, it might be a valuable learning experience to go flying with another, more seasoned pilot in the right seat. This is a great way to get “on the job training” without scaring yourself. Just be sure to obey two rules: Know and trust who you’re flying with and thoroughly brief who is pilot-in-command before starting the engine.

Make It Deliberate

All of these strategies demand careful planning and discipline; they can’t be used as shortcuts or excuses for poor decisions. But grappling with tricky decisions and expanding your personal skills envelope is one of the real joys of being a pilot.

They’re also an acknowledgement that one of the most powerful safety tools we have as private pilots is the ability to control when and how we fly. We aren’t beholden to chief pilots or customers and we can each make our own Standard Operating Procedures. If we give up that tool without so much as a second thought, we’ve made flying less useful, less safe or both.

Next time you’re planning a flight, don’t ask yourself “go or no-go?” Instead, consider “under what conditions would this flight be safe and enjoyable?” At the very least, it’s a valuable exercise for your decision-making muscles.


John Zimmerman is Vice President of Sporty’s Pilot Shop’s Catalog Division. He regularly flies a Citabria and a Pilatus PC-12. You can find John’s blog on Air Facts at airfactsjournal.com.

The post To Go, Or Not To Go? appeared first on Plane & Pilot Magazine.

]]>
Fly More This Winter https://www.planeandpilotmag.com/article/fly-more-this-winter/ Mon, 12 Dec 2016 21:15:41 +0000 http://www.planeandpilotmag.com/?post_type=article&p=25017 Don’t automatically lock your airplane in its hangar this winter. The cold months can be some of the best times to fly.

The post Fly More This Winter appeared first on Plane & Pilot Magazine.

]]>
Defining winter by the severity of cold weather on the North American continent can be a difficult task. In most years, anything south of a line through Atlanta, Little Rock, Oklahoma City, Santa Fe and Los Angeles has a good chance of a modest winter. Any location above 35 degrees north latitude can reasonably expect colder conditions. In this case, we’re not talking about the extremes of frigid temperatures, say, colder than -30 degrees F. It’s possible to operate in far more severe conditions—Alaskan bush pilots do it all the time—but we’re not considering those instances for the moment.

winter flying

Perhaps contrary to popular belief, winter weather is often friendlier to an airplane than the atmospherics of summer. For one thing, the flying can actually be better. In winter, the lifting forces that produce thunderstorms and tornadoes simply aren’t present. That’s not to suggest that winter storms can’t be dangerous, but with the exception of inflight icing and an occasional slick runway, pilots and their airplanes see mostly positive benefits from winter weather. No question, icing is a real hazard, but it’s not in the same class as CBs and tornadoes.

Unfortunately, humans are less impervious to the ills of winter than the airplane. For that reason, you’ll want to check and service the heater before your flight. Long before the advent of GPS, a friend was flying a well-used Shrike Commander he’d just purchased in Wichita back home to California and got trapped on top in very cold temperatures with a malfunctioning heater. The airplane was running fine, but the avionics weren’t. Joe was forced to stay up high for four hours with an inoperative heater before he finally cleared the undercast and was able to descend and land in New Mexico, a very cold and humbled pilot. He hadn’t checked the heater before takeoff, and his one good navcom had failed shortly after level.

Survival gear requirements change with the season, and you can spend as much as you wish to provide a hedge in case you need to land off-airport in frigid conditions. I fly a four-seat retractable that almost never carries anyone in the back seat, so I have all the room and weight allowance I need for emergency gear. On winter flights across rugged terrain, I have a habit of loading my Mooney with enough survival gear to keep a squad of Marines alive for two weeks at the North Pole. (My friend, Doug Ritter, of Equipped To Survive, offers a more realistic variety of information and gear for every survival need. Doug edits and publishes a newsletter specifically dedicated to all aspects of survival. Contact him at dougritter.com.)

winter flying
Flying in cold weather means that an off-airport landing even with no injuries can be a life-threatening event. Be sure to prepare with warm clothing and an emergency survival kit.

Winter brings with it fast-moving jet streams that can provide excellent tailwinds for those traveling east. “The Jet” was discovered in the mid-1940s when our B-29s, flying from Saipan and Tinian, first began saturation bombing of Japan. The ferocious westerly winds weren’t a new phenomenon, but they slowed our bombers by as much as 150 mph on the inbound run to Tokyo and other targets. At least the trip back to the Marianas was usually a quick one.

Here in North America, the jet stream migrates south out of Canada every winter, often bringing spectacular westerly winds to much of the United States. Those winds rarely descend much below 10,000 feet, but may still add an extra 20 to 30 knots in the bottom two miles of sky if you’re headed east. Pilots flying westbound, into the wind, can often mitigate the effect of headwinds by operating at the lowest possible height.

winter flying
One of the biggest winter risks is runway contamination. Snow- and ice-encrusted surfaces mean longer landings and dicier directional control. Photo by Shawn Roberts.

Back in March 1994, I took advantage of winter tailwinds to set eight new, city-to-city speed records between Los Angeles and Jacksonville, Florida, flying a new Mooney Bravo at FL250 all the way. The fastest leg was 338.32 mph, one hour and 59 minutes between LAX and Albuquerque. Total time coast-to-coast was seven hours and nine minutes, for an average speed of 300.2 mph over the 2146 sm course to Jacksonville. All those records still stand in class C1C.

The Bravo, predecessor to the current Acclaim, was/is the fastest piston, production airplane in the world, but such near-turboprop velocities wouldn’t have been possible without the help of the winter jet stream.

In addition to strong winds aloft, winter brings with it colder air (duh!), and that can be a major positive. The thick, dense air associated with winter temperatures is a wondrous thing for an airplane’s wing, prop and engine.

One slight negative of operating in cold temperatures is the occasional need for preheat. A standard start at 68 degrees F requires only about half the electrical power of one at 0 degrees F.

There’s little consensus on precisely when to preheat, but everything becomes easier when you bring the engine up to a reasonable temperature before engaging the starter.

Most experts on engine-operating techniques advise that it’s a false economy not to preheat if the overnight temperature fell below 20 degrees F the previous night. Most FBOs in northern locations do a steady diet of preheats in winter, and given an hour’s notice, they can have engine temps practically in the green when you show up and are ready to fly.

Another possible benefit of a preheat is that you may be able to warm up the cabin. Under some circumstances, you can ask the FBO to channel one of the heat ducts through the storm window to warm the cabin while the engine tubes are bringing cylinder head and oil temps off the bottom peg.

Icing in all its forms is a subject worthy of an entire book, in fact, dozens of books. For that reason, we’ll confine our examination of winter flying to problems on the ground and leave inflight icing for another time.

winter flying
Clear ice is a danger on the ground and in the air. Pilots must ensure that their planes are ice- and frost-free to avoid a tremendous loss of lift and potential catastrophe.

If you’re traveling and need to park the airplane overnight, you’ll need to decide between leaving it outside or parking inside a hangar. Obviously, a heated hangar is the preferred option. It’s also the most expensive one. In some places, you’ll pay $250 to $300 to house a single in a heated hangar overnight, $350 to $400 for a twin.

A cheaper alternative may be to have the airplane pulled inside for perhaps two hours just before departure, long enough to bring everything up to a reasonable temperature. That may not be much cheaper than leaving it inside all night, since the big fee is for the power to reheat the hangar rather than the occupied floor space.

If that’s not an option, you can leave the airplane outside and take your chances on frost. Should you step out in the morning and find your Piper has been transformed into a popsicle, all may not be lost.

A full deice with type IV ethylene glycol is usually fairly expensive, and that isn’t always available unless you can move the airplane to a specific deice pad, plumbed to drain away deice fluid safely.

Some pilots carry a few gallons of antifreeze with them, but that’s a bad idea. Antifreeze is sickeningly sweet and will kill any animal that licks it up. Additionally, most airports frown on pilots pouring anti-freeze on the wings and tail, inconveniently leaking some on the ground. As you might imagine, antifreeze and ground water combine into toxins you don’t want to think about.

The next option is to use a standard credit card to remove frost and snow. I have a long-outdated Long Beach Airport access card that’s thick and strong, and will blade away most snow and frost fairly quickly without shattering or damaging the paint. A standard American Express card may also work, but you stand a good chance of breaking it and having to rely on cash for the rest of your trip.

Once the airplane is clear of ice and it’s time to depart, winter may offer you too much of a good thing. You may develop a density altitude problem that’s exactly the opposite of what you might imagine.

Cold starts should obviously follow your airplane’s start procedure, but a universal rule is not to over-prime. Using too much prime can wash the cylinder walls clean of any residual oil and result in metal-on-metal contact when you engage the starter. The jury is still out on pulling the prop through before initiating start. Personally, I do it on any cold-weather start.

Everyone who operates an airplane in mountainous terrain understands the concept of high-density altitude. Fly into Leadville, Colorado, in summer, and the temperature can seem very comfortable at, say, 50 degrees F. The “yeah, but!” comes when you start calculating what the temperature “should be” at Leadville’s 9,927 feet MSL elevation. Standard temp at that height works out to roughly 29 degrees F. If the actual temperature is 50 degrees F, the air is considerably thinner than standard, equivalent to about a 12,000-foot density altitude.

winter flying
There are advantages to winter flying, including enjoying low density altitudes, even when you’re operating in the High Country.

Nonstandard temperatures can cut both ways, however. Pilots address this problem on a regular basis in the Far North at more typical, near-sea-level airports when the winter temp drops to -40 degrees F or colder. Since standard temperature should be 59 degrees F, they’ll be operating at about 100 degrees F below normal temperature. That produces a density altitude of almost -7,000 feet (note the minus sign). That represents a kind of natural supercharging.

Now, see what happens when you translate that density altitude to engine power. A perfectly tuned, normally aspirated piston engine, operating in sea-level, standard conditions, should deliver something like 28.5 to 29.0 inches of manifold pressure. Ideally, induction systems should produce 29.90 inches, but no induction system is perfectly efficient. Most lose at least an inch of pressure to inefficiencies in the system.

Go to full power at conditions that simulate 7,000 feet below sea level, and your engine will suddenly deliver more like 36 inches MP at full power. For most engines, that means you’re technically over-boosting the engine.

If you monitor the manifold pressure and are careful to maintain it at 30 inches or less, winter temperatures can allow you to enjoy slightly more power and performance without the risk of engine damage.

winter flying
For pilots used to operating in warmer climes, the concern is usually high oil temps. In cold weather, the opposite is true and low oil temps are a real hazard to safe and reliable operations.

That much power might feel great under your right palm, but it could also be destructive if maintained for more than a few minutes.

Conversely, such dense air is wonderful news for most airfoils. Wings love a compressed sky, and propellers produce more thrust as the air becomes thicker and they can take a bigger bite with each passing blade.

Climb also benefits from cold weather, since there’s little convective activity. The ride uphill is usually smooth and quicker than normal. Traditional wisdom suggests cruising higher than you normally do if there are tailwinds available.

In winter, visibility is typically better than during the haze of summer when you’re not fighting a snowstorm. In fact, much of the time, you’ll be greeted by clear air and sunshine, even if it’s a little chilly.

On a recent Columbia 400 delivery to Geneva, Switzerland, I encountered something I’d never seen before: an oil temperature of what I interpolated as 50 degrees F, well off the bottom of the gauge. (I say “interpolated” because there were no markings below 135 degrees F.) Flying out of Bangor, Maine, for Goose Bay with an outside air temperature of -30 degrees F, I first assumed it was simply a malfunctioning oil temp gauge. Everything was running well, the weather was severe clear and the engine didn’t seem unhappy. I continued to Goose Bay, landed and immediately taxied to the maintenance shop. The mechanic said he’d seen that before, and it wasn’t unusual when conditions were what he termed “brisk.”

Trouble was, my next three legs were across the North Atlantic, and I wasn’t about to contend with a possible engine failure in such “brisk” conditions.

A few phone calls later, I determined that such amazingly low readings weren’t that unusual in northeastern Canada in winter, and the Columbia 300/350/400 had experienced the problem before. The Goose Bay mechanic installed some metal deflectors that partially blocked the cooling air intake, and the oil temperature climbed to about 135 degrees F on my next leg, warm enough to justify continuing the trip.

Winter takeoffs and landings can also present some special problems you won’t encounter in summer. If there’s snow on the runway, most any idiot can figure there might be ice underneath, but if the runway looks clear!.

In reality, you may be better off when there’s visible snow available for taxi and takeoff. In some parts of the Far North, it’s not always possible to keep the taxiways and runways clear because of the rate of snow accretion. For that reason, you’ll often taxi by reading the guide poles that rise a foot or more above the taxiway/runway edges to help delineate aircraft movement areas.

winter flying
Icing forecasts are often guesses, at best, so smart pilots prepare for the possibility of delays and always leave themselves an out.

Sometimes, even that isn’t enough. A few years back, I was delivering a 58P Baron to Cleveland Hopkins following a double-engine overhaul by Victor Aviation in Palo Alto, California. The snow at Cleveland was heavy, but the air was characteristically smooth when I arrived at 9:00 p.m.

The Baron made the approach look easy as I broke out near minimums, touched down in heavy snow and almost immediately lost all forward visibility to low-lying mist and snow. With plenty of runway ahead, I stayed off the brakes and let the airplane roll out straight ahead. After I’d stopped, the tower called and said they couldn’t see me, only fair since I couldn’t see them, either.

The controller started sending everyone around, but it barely mattered as the airport had dropped below minimums. They advised me to hold position, turn on every light in the airplane and wait for a follow-me truck.

It took about 10 minutes before the truck found me, starting from the opposite end of the runway and driving very slowly toward me. He came up on ground and told me to stay right behind him as we taxied to the ramp at a slow walk.

It seemed to take forever until a huge hangar with an open door and welcome floodlights materialized in the snow. Fortunately, Cleveland was my destination, so I didn’t need to contend with snow on ice the next day.

The point is, don’t trust snow covering a runway or taxiway on the premise that there’s no ice underneath. You may not know it’s there, but if it is, the airplane will find it for you.

Understandably, airports always clean their runways first, and taxiways/ramps are a second priority. Keep taxi speeds as slow as practical, and don’t automatically try to take high-speed turnoffs just because you think you can. Make all turns as gently as possible, and stay off the brakes, especially if you’re unfamiliar with the airport layout or if visibility is marginal.

Even what looks to be a clean runway can be deceptive. Many years ago, during a Christmas flight from Long Beach, California, to Tampa, Florida, I landed at perpetually windy Amarillo, Texas, with a slight crosswind and what looked to be a clear runway, only to discover it was covered with black ice.

My tailwheel Bellanca Cruisemaster started a slow turn to the left and kept right on going through a full 360, tracking the centerline and ignoring whatever I was doing wrong with rudder and brakes. It finally stopped, once again pointed down the centerline just as if I’d planned it that way. I hadn’t.

As I sat there with my heart still racing and the airplane totally undamaged, the droll controller felt compelled to ask, “Bellanca 85 November, do you require assistance to taxi?”

As of January 1, 2016, Senior Editor Bill Cox has logged 15,100 flight hours in 321 types of aircraft.
He also holds 28 world city-to-city speed records, has made 211 international delivery flights, and owns
and flies a LoPresti Mooney. You can email Bill at flybillcox@aol.com.

Check out these 20 Winter Flying Tips to fly more during the season.

The post Fly More This Winter appeared first on Plane & Pilot Magazine.

]]>
A Rapidly Closing Weather Window https://www.planeandpilotmag.com/article/rapidly-closing-weather-window/ Wed, 05 Oct 2016 11:36:09 +0000 http://www.planeandpilotmag.com/?post_type=article&p=24224 The pilot of a JetPROP-converted Piper Malibu tried to thread his way through severe weather without airborne radar

The post A Rapidly Closing Weather Window appeared first on Plane & Pilot Magazine.

]]>
It happened on June 18, 2014: With thunderstorms popping, the pilot of a turbine-powered Piper Malibu PA46-310P JetPROP conversion seemed to be doing a good job of weather avoidance, but then made a turn and flew into a monstrous cell. The airplane crashed at Lehman, Texas, killing the pilot and both passengers. It took two years and four days for the NTSB and its investigators to figure out as much as they could about what happened, and issue their report and probable cause statement. The Safety Board said the accident was due to “the pilot’s improper decision to enter an area of known adverse weather, which resulted in the loss of airplane control. Contributing to the accident was the air traffic controller’s failure to provide critical weather information to the pilot to help him avoid the storm, as required by Federal Aviation Administration regulations.” While it’s difficult to argue with the probable cause, it only reflects part of the story. For example, you won’t find it in the NTSB’s report, but it turns out that the airplane wasn’t equipped with airborne weather radar. A contact at the NTSB was kind enough to dig out that information when I asked about it. And, as you’ll see, there’s more.

The instrument-rated private pilot had logged 2,258 hours with 188 in the Malibu conversion. His third class medical required glasses for distant and near vision. He told the FAA he wasn’t using medications, but toxicological testing after the accident found he was using the erectile dysfunction drug Sildenafil, the sleep aid Zolpidem, the pain reliever Ibuprofen and the heartburn medication Ranitidine. He also was found to have used marijuana. The NTSB said that it’s not likely the drugs contributed to the accident, but can anyone be sure there wasn’t even a minor cumulative effect? Investigators couldn’t determine from his flight records whether the pilot had a current flight review and was current for instruments.

The flight was from Aspen-Pitkin County Airport/Sardy Field (KASE) at Aspen, Colorado, to Brenham Municipal Airport (11R) in Texas, a direct distance of 851 statute miles. It was supposed to take 3 hours and 30 minutes. The pilot had filed a cruise speed of 250 knots at FL270 (approximately 27,000 feet). The pressurized six-seat PA46-310P Malibu typically can get up to FL250 where it cruises at about 200 knots. This airplane, which was built in 1985, received an STC conversion to a much more powerful 560 shp turbine powerplant in 1999. With the Pratt & Whitney PT6A-34 engine, it could fly higher, faster and get above more weather. The NTSB pointed out a minor complication, which the pilot ignored, and really didn’t contribute to the accident. Although the Pilot Operating Handbook for the turbine conversion approved operation as high as FL270, the airplane’s altimeters were only certified to 25,000 feet.

At 1:09 p.m., Mountain Daylight Time, about 17 minutes before departure, the pilot phoned flight service for a weather update and to file his instrument flight plan. He indicated that he already knew some of what was going on with the weather, and was given an abbreviated weather briefing. The briefer described the flow of moisture north from the Gulf of Mexico, and discussed a Convective SIGMET for a developing line of thunderstorms near the destination. If, indeed, the pilot had given himself a comprehensive weather briefing on the computer before telephoning flight service, he would have been ready for a mix of visual and instrument conditions en route. There would be some turbulence getting out of the Aspen area, but then a fairly good ride with the wind at FL270 from 240 degrees at about 32 knots. The airplane could be in and out of the clouds, especially once over Texas. With all the moisture and cloud buildups, and an outside air temperature of minus 25 degrees C, the pilot might have to give the airplane’s de-ice boots a bit of a workout.

By 3:17 p.m., Central Daylight Time, after having been airborne for about 51 minutes, the pilot was being handled by Albuquerque Center at his cruise altitude of FL270. When the pilot checked in, he reported a smooth ride. The controller announced that a Fort Worth Center Weather Advisory could be heard on the flight watch frequency. A few minutes later, the Malibu was handed off to another Albuquerque Center controller. At about 3:49 p.m., this controller announced that a Convective SIGMET for Texas had been issued and could be heard on flight watch. Neither of the controllers broadcast details of the unfolding weather. Pilots tuning to flight watch would have heard that severe thunderstorms had developed and were forecast to continue over portions of Oklahoma and Texas. Hail was forecast to be up to 2 inches, there was extreme turbulence, surface winds could be expected to hit 60 knots, and the cloud tops could reach up to 55,000 feet. Along the route of the Malibu’s flight, it might not be quite so bad, but it was still severe, with cloud tops to 45,000 feet, hail to 1½ inches in diameter and wind gusts on the ground to 50 knots.

At 3:54:17, the controller alerted the Malibu pilot that “…I’ve seen you as much as five hundred feet high and as much as four hundred feet low in the last minute.” The pilot replied, “….we’re having some autopilot issues here on the hold, uh, we’re trying to get her…” The rest was unintelligible.

Whether the airplane was being bounced around in heavy updrafts and downdrafts, or really did have an autopilot problem, wasn’t pursued at the time. After the accident, the autopilot was checked by the manufacturer and seemed to be okay. About two minutes later, the pilot radioed, “…we’re going to do ten degrees left for build-up.” The controller said, “…left deviation approved and, uh, advise on course.” The pilot confirmed, “…we’ll advise back on course.” Instead of turning to the left to fly east of his course, the pilot turned to the right, deviating west of course.

Over the next few minutes, the controller’s scope continued to show thunderstorms popping, and the Malibu pilot continued his maneuvering for weather avoidance. At about 4:06, the controller radioed, “…does it look like you’re going to, ah, to the west of that cell that’s off your, ah, twelve o’clock all the way down to your, ah, nine o’clock position, you going west of that…” The pilot responded, “Yeah, we’re going west of that one and we might be able to go between those, we’ll take a look and see, but otherwise we’ll go around it.” The controller radioed,
“…not a problem, uh, deviations right of course are approved and, uh, just advise when you can make it to your, ah, destination.” The controller put the code “DR” in the data block adjacent to the Malibu’s radar return indicating that the airplane was deviating to the right. The flight was then handed off to Fort Worth Center.

When the pilot checked in with the new controller, he advised being at FL260, but was going back up to FL270 while doing a turn for weather avoidance. The controller acknowledged, and didn’t volunteer any information about the weather she was seeing on her scope. After all, the indication “DR” meant he was deviating and, therefore, had to know about the weather. She did ask if the pilot was going to go north of a cell. The pilot responded, “…ah, we’re trying to go in this window here.” The controller asked the pilot to “…say again…,” and he told her, “We’re trying to go through a window here.” The controller replied “…right, thanks…” and, about 30 seconds later, asked whether the pilot needed FL290. She asked twice more whether he needed flight level 290. At about 4:32, she radioed, “…I show you’re right in the middle of moderate extreme precipitation, did you need a different altitude?” About 20 seconds after that transmission, the accident pilot transmitted, “Mayday.”

“Thunderstorms popping… the Malibu pilot continued his maneuvering for weather avoidance. …[ATC queried about storm clearance], and the pilot responded, “Yeah, we’re going to be west of that one and might be able to go between those…we’ll take a look and see…”

When the controller didn’t immediately respond to the mayday call, the pilot of an Eclipse jet operating at FL410 relayed to the controller that there had been a mayday call. He also heard the pilot say that he was “spinning.” The controller radioed the Malibu, asking for the pilot’s altitude. The pilot replied, “Nineteen.” A United Airlines pilot reported hearing the Malibu pilot say he was spinning and had lost sight of the ground or horizon. The controller asked American Airlines and Alaska Airlines pilots to try to make contact.

FAA radar data, as captured by a commercial service, showed that just before the mayday call, the airplane had entered a climb of more than 1,000 feet per minute. About a minute later, it went into a descent of almost 3,000 feet per minute. Less than a minute after that, the descent reached a rate of 6,774 feet per minute, then slowed to 4,500 feet per minute. The airplane was descending at a bit under 3,000 feet per minute just before radar contact was lost.

The NTSB said the crash occurred at 4:35 p.m., in an open field about one-half mile east of Lehman, Texas. The wreckage was pretty much intact, and showed extensive vertical crushing. Investigators concluded that the airplane had come down in a flat spin.

An NTSB meteorologist overlaid the Malibu’s flight track on weather radar images and lightning strike data. Beginning at about 4:00, the pilot would have flown between cells reaching 27,500 and 34,000 feet. By 4:12 p.m., cells to the east and south had grown to 44,000 feet. Within the next four minutes, storms to the south had grown to 47,000 feet. Some cells within 10 miles of the airplane contained hail greater than 2 inches in diameter. At 4:33 p.m., instead of continuing to avoid the storm cells, the airplane turned to the east-northeast into a growing area of intense activity. An intense core was only seven miles east of the airplane’s location. The cell was 20 miles wide and reached about 46,000 feet. The pilot flew directly into it.

The NTSB’s report and published backup material don’t document the avionics carried onboard the Malibu. One might expect to find airborne radar on a Malibu like this one, especially when the pilot plans to operate in the flight levels during thunderstorm season, but not all have it. Some Malibus have a radar set that uses a small antenna located in a wing. Others use an aftermarket higher-powered set with a larger, more effective antenna in a pod hanging from the right wing.

My NTSB source was told by an investigator that the airplane was equipped to receive and display weather data. The investigator said they couldn’t determine whether the pilot had a current subscription to a weather service, nor what the pilot may have been viewing during the accident flight. If he was still in visual conditions, the pilot should have been able to see that he was headed into some towering giants. If he was looking at a NEXRAD mosaic, any window he saw between cells could already have closed. There are processing and transmission delays when NEXRAD images are sent by commercial services. Even though there are time hacks associated with the images, sometimes pilots will fail to make the connection. The NTSB, for one, has warned about this potential hazard. We don’t know the level of the pilot’s training in interpreting weather radar images. If he was looking at a radar image, old or real-time, was his window through the weather something he misinterpreted, or really in the image as a result of signal attenuation? One thing we can do to help avoid repeating what happened here is to learn more about reading radar.

In looking at wreckage photos, I could see that the airplane’s panel did have an Argus lightning detector. If it was working, the pilot might have seen some of the 36 cloud-to-ground and 2,117 in-cloud lightning strikes recorded within 25 miles of the accident site at around the time of the crash. A lot of the lightning was in the direction he had turned, and should have been taken as a warning to stay away.

The NTSB report doesn’t provide much enlightenment about how the pilot wound up in a spin. We might speculate that the pilot entered clouds and there was some ice accumulation, resulting in decreased aerodynamic capability. Couple that with the airplane at its maximum approved altitude, perhaps forced even higher by updrafts, throw in varying angles of attack due to turbulence, and you have an aircraft perilously close to its operational threshold. A military pilot near the Malibu, but lower, reported picking up ice. Testing after the accident found the Malibu’s de-ice boots were functional.

The Eclipse pilot I mentioned earlier was bothered by what sounded to him like the controller’s laid-back or laissez-faire reaction to the unfolding emergency. He subsequently telephoned ATC to discuss it, and was assured that everything was handled according to current directives. He reported that his Eclipse was equipped with X-band airborne weather radar, a StormScope and XM Satellite Weather. He said at FL410 he had to deviate around the line of weather that was over the accident location. He estimated the cloud tops were at 48,000 feet and there was a lenticular-shaped cap cloud on top. He said he briefly found himself in instrument conditions while maneuvering around the weather. He said that the weather was very deceptive around the time of the accident, with cells growing rapidly and merging. He believed that the accident airplane, down lower, would have been in IMC.

While the NTSB faults the controller for failing to volunteer weather avoidance information, we have to remember that the pilot didn’t ask for any kind of help until it was likely too late. A simple request to the controller a bit earlier for vectors or a request from the pilot for her to describe the weather returns she was seeing might have resulted in a different outcome. PP

Peter Katz is editor and publisher of NTSB Reporter, an independent monthly update on aircraft
accident investigations and other news concerning the National Transportation Safety Board.
To subscribe, visit www.ntsbreporter.us or write to: NTSB Reporter, Subscription Dept., P.O. Box 831, White Plains, NY 10602-0831.

Want to read more proficiency analysis from Plane & Pilot? Visit our After The Accident archive.

The post A Rapidly Closing Weather Window appeared first on Plane & Pilot Magazine.

]]>
Choosing VFR Cruise Altitude https://www.planeandpilotmag.com/article/choosing-vfr-cruise-altitude/ Thu, 22 Sep 2016 19:25:49 +0000 http://www.planeandpilotmag.com/?post_type=article&p=23907 How’s the ride up there? Follow these 5 simple steps for smoother flying

The post Choosing VFR Cruise Altitude appeared first on Plane & Pilot Magazine.

]]>
They say the three most useless things to a pilot are runway behind you, fuel not in your tanks, and altitude above you. So when you’re choosing your VFR cruise altitude for your next cross-country, is higher really better? It could be, but you have a lot to consider. Here are five things to think about when you’re planning your next flight.

VFR

Tip 1: Am I Going To Hit Something?

There’s nothing that will ruin your day like hitting terrain or an obstacle. So how do you make sure you’re clear on your route? If you’re flying VFR, one of the easiest ways is to open your sectional and check out the MEF (Maximum Elevation Figure) altitudes for your route.

VFR
If you’re flying VFR, it’s as easy as opening your sectional and checking the Maximum Elevation Figure altitudes for your flight.

The MEF is the bold blue altitude, in hundreds of feet MSL, listed in the middle of each quadrant of your sectional. That altitude guarantees you at least 100 feet (up to 300 feet, in some cases) of clearance from all terrain and obstacles in the quadrant.

So, as long as you pick an altitude above the MEF, you can rest easy in knowing that you’re not going to hit something poking out of the ground while you’re en route.

Tip 2: Can My Plane Actually Get There?

It’s a valid question, and it’s not always easy to answer. You need to be practical with your altitude choice. If you’re flying a short distance, it doesn’t make a lot of sense to spend the majority of your flight in a climb.

VFR
Be practical with your altitude choice, and use your aircraft’s Fuel, Time and Distance to chart.

That’s where your aircraft’s Fuel, Time and Distance to Climb chart comes into play. For most aircraft, your time-to-climb is pretty linear, but if you’re flying a normally aspirated airplane above 10,000 feet MSL, your climb rate can start to tail off significantly. And, on top of that, you’re burning extra fuel, and flying a slow indicated airspeed, all the way to your cruise altitude.

But the opposite is true when it comes to your true airspeed. The higher you go, the higher your true airspeed. The rule of thumb is that you gain 2% of true airspeed for every 1,000 feet you climb, and that can make a big difference. Consider this: If you’re flying at 140 knots indicated at 5,500 feet MSL, your true airspeed will be roughly 154 knots. But if you fly the same indicated speed at 11,500 feet, your true airspeed shoots up to 170 knots. That’s a gain of 16 knots, which is a big difference maker, especially on long flights.

Tip 3: What Kind Of Airspace Do I Need To Deal With?

Ah, everyone’s favorite: airspace. There’s controlled airspace, special-use airspace and just about every kind of airspace you can think of listed on sectional charts these days.

VFR
Flight-planning tools can help you navigate around tower-controlled airports and special-use airspace; when in doubt, just fly above it!

Fortunately, there are lots of great planning tools out there, like ForeFlight, which can help you navigate around tower-controlled airports and special-use airspace along your route. But there’s another way to make life easy on yourself when it comes to airspace: Simply climb above it.

If you can get yourself above 10,000 feet MSL, you’ve all but guaranteed yourself clearance above tower-controlled airspace, even Class B. There are, of course, a few exceptions, like the Denver Class B that extends up to 12,000 feet MSL, but they’re few and far between.

Unfortunately, the same can’t be said for restricted areas and other special-use airspace, but a quick check on your sectional map can clear up any questions about that.

Tip 4: What About The Clouds?

Now that you’ve gotten this far, you need to contend with the weather. And Mother Nature isn’t always cooperative when it comes to flying.

VFR
Use your METARs, TAFs and PIREPs. You may be able to climb above few or scattered clouds; if you’re looking at a broken layer along your route, it’s best to stay below.

That’s where your METARs, TAFs and PIREPs come into play. When you’re checking the clouds, think about coverage and altitude. If you’re looking at few or scattered clouds, climbing above them might be an option, but if you’re looking at a broken layer along your route, it’s best to stay below.

After all, there’s nothing more embarrassing (and panic-inducing) than getting stuck on top of a cloud deck with no way to get down, short of declaring an emergency, or, if you’re an instrument pilot, scrambling to find charts to navigate your way down through the soup on a pop-up IFR clearance.

Also, remember that METARs and TAFs list cloud bases in AGL, not MSL. So you’ll need to do some math to figure out where the bases will be to maintain your VFR cloud-clearance requirements.

Tip 5: Are My Passengers Going To Hate Me?

There’s one final consideration, and it’s quite possibly the most important thing: What are your passengers going to think of you when you touch down after your flight?

VFR
Everyone hates turbulence. Be on the lookout for “smooth ride” altitudes, as well as the favorable winds aloft.

If your passengers’ teeth are getting rattled out of their heads because of turbulence, they’re not going to be very impressed. And one place you’re almost guaranteed to find turbulence is around shear layers in the winds aloft.

While you obviously want to consider your headwind or tailwind along your route, you also want to make sure you’re keeping yourself clear of any significant shear layers aloft.

Take this, for example. On this route from KGCY-KEHO, there’s a 24-knot wind velocity difference between 3,000 feet and 6,000 feet, with a nearly 50 degree wind direction difference. And, if you’re thinking things would be bumpy in that area, you’re right.

Taking a look at area PIREPs with a tool like ForeFlight (or ADDS, if you’re not a ForeFlight user) confirms what you’d expect: A Cherokee pilot reported continuous moderate turbulence below 4,500 feet MSL.

Unless you want to pack extra sick sacks for your passengers, be on the lookout for the “smooth ride” altitudes, as well as the favorable winds aloft.

Putting It All Together

There’s a lot to consider when you’re picking your cruise altitude. But if you’re thinking about obstacles, your plane’s performance, and the weather and winds along your route, you’ll have a smooth flight, and, hopefully, some happy passengers, as well.

Colin Cutler is a Boldmethod co-founder, pilot and graphic artist. He’s been a flight instructor at the University of North Dakota and an airline pilot on the CRJ-200, and has directed development of numerous commercial and military training systems. Visit the Boldmethod.

For more flying tips, check out our “Winds Aloft” article for finding a friendly push!

The post Choosing VFR Cruise Altitude appeared first on Plane & Pilot Magazine.

]]>
Winds Aloft: How To Find The Occasional Friendly Push https://www.planeandpilotmag.com/article/winds-aloft-find-occasional-friendly-push/ Mon, 01 Feb 2016 17:24:24 +0000 http://www.planeandpilotmag.com/?post_type=article&p=20538 But use caution: the wind might be a little tricky

The post Winds Aloft: How To Find The Occasional Friendly Push appeared first on Plane & Pilot Magazine.

]]>
WindSock

Most prospective aviators are excited about joining Lindberg, Yeager and Hoover in the sky, but they’re usually less enthusiastic about investigating the ways of weather, at least until they start flying places. In this case when we say “weather,” we’re talking mostly wind. For safety’s sake, we need to understand the movement of fronts, development of thunderstorms, the hazards of fog and ice, and the myriad of other weather phenomenon about which meteorologists wax poetic, but even if conditions are CAVU from point of departure to destination, winds affect almost every flight.

The good news is that it’s easy to find wind and winds aloft information these days, whether it’s from an official briefer, the internet, a handy app or even from the TV.

We don’t necessarily need to become diagnosticians of winds in order to understand how to turn weather information to our advantage. Pilots flying short hops on Sunday mornings for pancakes or burgers may not need to do more than look out the window to make a reasonable go/no go decision, but aviators planning cross-country flights to far horizons need to understand what they’ll be getting into well before they get into it.

That need becomes especially acute when the distance extends from 50 or 100 miles to 500 nm or more. A more comprehensive forecast becomes essential when the trip exceeds known parameters, ranging from across town or state to across a continent.

That’s because individual weather fronts are usually confined to a comparatively small area. Extend your reach to 1,000 nm, and you may fly through several systems with winds blowing in different directions. Flight planning over long distances requires a careful analysis of winds aloft.

The uninitiated might expect the laws of probability to produce headwinds and tailwinds with roughly the same frequency, effectively canceling each other. Not so.

Fly a one-way trip in a consistent direction over 300 nm or less, and variable winds aloft may not present much of a problem, but if you’re planning an out-and-back over any distance or a hop longer than 500 nm, you’d best be aware of winds at every point and every altitude along your route.

The uninitiated might expect the laws of probability to produce headwinds and tailwinds with roughly the same frequency, effectively canceling each other. Not so. In the Northern Hemisphere, wind patterns are predominately west to east. That might suggest you’ll usually break even if you’re flying a semi-horizontal track above the equator.

As many of us learned in flight school, however, even a direct tailwind will be more than offset by a direct headwind on the return leg. That’s because a headwind acts on the airplane longer than does a tailwind.

A simple example is a 100-knot airplane flying into a 20-knot headwind on a 200 nm trip. Time enroute will be 2+30 for such a trip. The return leg, with a 20-knot tailwind, will demand only 1+40. That’s a total of 4+10. In no-wind conditions, the same airplane would need only four hours for the round trip. No matter what combination of wind speed, wind direction and aircraft speed you choose, the result is always the same. As long as all other factors remain constant, you lose.

B_29

JET STREAMS
Phenomenal jet streams in the high sky were first discovered by WWII B-29 bomber crews in the mid-1940s on the way from Tinian, Marianas, to Japan. Winds were reported as high as 100 knots on some of those missions. Here at home, the reality is closer to 50 knots or less above FL300. If corporate or airline jets can pick up even 10 minutes on a four- or five-hour trip, however, that represents a major fuel savings, not to mention a reduction of crew duty time. (Airliners can burn huge amounts of fuel; just under a gallon a second on a 747 at cruise. In practical terms, a 10-minute reduction in flight time in a 747 results in roughly 600 gallons less fuel burn.)

Any consistent wind on a round- trip flight will always reduce your net groundspeed. That’s because any tailwind will never save as much time as a headwind will lose. Even a direct crosswind will resolve to a headwind component. In other words, you’re actually better off with no wind at all, rather than any consistent wind.

Winds are a big deal to those of us who own or operate aircraft in the speed regime below 160-170 knots, where wind factors can represent a great percentage of our cruise speed. A 10-knot wind comprises fully 10% of the cruise speed of a Cessna 150 but only five percent of a Malibu’s cross-country velocity. (That said, wind considerations at the flight levels can be critical as well, as very high winds aloft can add greatly to the time it takes to fly a westbound leg.)

Run a few test situations through your electronic E6B, and you’ll find that you won’t begin to realize any significant benefit from tailwinds unless they blow from the aft 160 degrees. The forward 200 degrees, including 10 degrees on each side from the aft quadrants, are universal losers or break even at best.

Of course, smart pilots might be able to cut their losses by flying as high as possible on any leg with a push; then, cruise lower operating into a headwind on the return. Another possible method of reducing total time enroute, at least for those strange folks who buy a fast airplane to fly slow, is to increase power on a headwind leg to reduce the total time enroute. This isn’t terribly efficient in terms of fuel burn, but it can help you make up some of the time lost to winds.

Just as with life, airline fares and the IRS, no one ever promised winds would be fair. The airlines appreciate this phenomenon better than anyone, and they’ll sometimes fly pressure patterns, i.e. displace course lines as necessary to realize better winds aloft.

If the high- and low-pressure systems are lined up properly, you may be able to deliberately route a short distance left or right of a great circle or other prescribed route to take advantage of tailwinds associated with the clockwise flow of a high pressure system or the counter-clockwise flow of a low. After all, time, not distance, is the only number that matters in efficient cross-country travel. (Is that the reason the airlines sometimes route through Dallas to fly from Los Angeles to Chicago?)

Pressure pattern flying is obviously most advantageous to high-speed aircraft operating in the upper flight levels. It also can be beneficial to general aviation in the middle altitudes and bottom of positive control, Type A airspace. I rarely subscribe to pressure pattern flying, as I’m usually operating non-RVSM-approved aircraft below 250 knots. By definition, I’m at or below FL280, so I’m rarely exposed to the super winds of tall altitudes.

There are a few other tricks to wind management that some pilots use. One is to always think “right” when approaching a weather system that you’d rather fly around than through or under. By definition, weather usually brings with it low pressure, and as mentioned above, that means counter-clockwise flow. If you need to circumnavigate weather rather than charge through it, always favor the right side if terrain permits, as that’s more likely the tailwind direction.

Also, while flying below a broken layer, take a tip from glider pilots and amend your route very slightly to fly beneath the major buildups rather than aim for the patches of clear sky. When I was training for the sailplane rating many years ago, I learned that glider enthusiasts call these clear sections “blue holes” and know they signify areas where the sky truly is falling.

In other words, you’ll very likely encounter downdrafts beneath the blue holes and experience updrafts under the major cloud formations. You should be better able to maintain cruise speed if you’re not fighting downdrafts, another factor that can make it more efficient to reach your destination.

wind-cruel

WIND IS CRUEL
Life isn’t fair. Here’s proof. This graphic shows the same plane going on an out-and- back trip with identical winds. While the trip out gets the benefit of a nice push, the punishment on the return trip makes it hardly worthwhile. Winds just don’t balance out. As the graphic shows, at 1,000 knots true airspeed with a windspeed of, likewise, a breezy 1,000 knots, a trip of 1,000 nm will take a mere half hour (1,000 nm/2,000 knots ground speed). Nice. The return trip won’t be as pleasant. With 1,000 knots on the nose, our sad pilot will never get home (1,000 nm/at zero ground speed—hovering until the gas runs out). We’d suggest turning around. Paris is nice.

Wind aloft forecasting would be considerably simpler if all you had to deal with was consistent west-to-east flow. Unfortunately, the diabolical weather gods weren’t about to make it that easy. Topography and Coriolis effect induce winds to perform sometimes weird tricks that you may not be able to predict.

In my home state of Alaska, Mt. Denali and Mt. Foraker, huge monsters of granite and ice located near the geographic center of the state, disrupt and deflect “normal” west to east weather flow. These sometimes confuse weather patterns in Fairbanks, 100 miles north, and Anchorage, 130 miles south. The mountains are infamous for making their own weather, molding the winds to fit their contour, reshaping the sky from its normal horizontal pattern to a decidedly south/north flow.

Years ago, on an amazingly clear day, with Anchorage Center’s permission, I flew a Cheyenne III across the rounded dome of McKinley (recently renamed Denali) at 20,400 feet, 80 feet above the peak, all for no better reason than I could. The wind was atypically dead calm, so the mountain apparently didn’t mind my indignity. As I turned back toward Point Barrow and climbed back up to FL270, Center asked about the ride, then, popped my bubble by advising a low level McKinley overflight was a common request on exceedingly rare, windless days.

Coriolis effect, the spin of the Earth, also tends to bend the sky from longitude to latitude. In the northern hemisphere, Coriolis turns weather systems stubbornly toward the Pole, often with amazing strength above 30 degrees north. In Winter, this can result in extreme southerly winds, and unless you can stop the Earth’s rotation, there’s little you can do but go with the flow.

Managing winds can be a game of inches, however. You can help maximize efficiency by managing climbs and descents more intelligently. If the breeze is friendly up high and the cylinder head temps are willing, climb at Vy or perhaps Vy plus 10 to reach the favorable winds as quickly as possible. Levitating into headwinds, I’ll fly a cruise climb at Vy plus 20 or 25 to maximize distance traveled for time spent.

Similarly, at the opposite end of a tailwind trip, I’ll maintain altitude as long as possible, then descend at a good clip with speed brakes deployed to destination.

I didn’t major in psychology, but I have a theory that another reason we see more headwinds than tailwinds is that we simply tend to remember them more readily. I like to analogize the art of psyching out winds aloft to what I call Murphy’s Law of Freeway Lane Choice; Whatever lane you’re in will always be the slowest. Sadly, the same disappointment often awaits us pilots when we arrive at that altitude we guessed would have the most favorable winds. Happily, like ill-fated lane-gambling motorists, we can always try a different altitude, too.

The post Winds Aloft: How To Find The Occasional Friendly Push appeared first on Plane & Pilot Magazine.

]]>
Learning To Love (Or At Least Accept) Winter https://www.planeandpilotmag.com/article/learning-to-love-or-at-least-accept-winter/ Mon, 28 Dec 2015 04:00:00 +0000 http://planepilotdev.wpengine.com/article/learning-to-love-or-at-least-accept-winter Flying is for all seasons

The post Learning To Love (Or At Least Accept) Winter appeared first on Plane & Pilot Magazine.

]]>
lead-2

Practically every spring brings proclamations that “flying season is just around the corner.” It seems many pilots regard late spring, summer and fall as the only time to fly. Some aviators in the Northern U.S. apparently prefer to lock their airplane in the hangar each October and not come back until late April.

That’s hard to understand, as winter has many things going for it. Summer heat spawns such often-unpredictable (and unflyable) phenomena as tornados, hurricanes and thunderstorms, whereas winter usually features nothing worse than snow, wind and cold temperatures.

I grew up in Alaska, where many residents regard winter as more of a blessing than a curse, a time for skiing, dog-
sledding, snowshoeing, ice skating and yes, flying. Up north, aircraft owners think nothing of warming up the Bonanza, Warrior, 150 or 185 for a flight whenever the weather is decent, no matter what the season.

It’s true flying in winter demands more work than at other times of year, but the enhanced performance, often spectacularly improved visibility and potentially greater range can make winter an equally favored season.

During 10 years in America’s largest state, I grew to enjoy the cold weather of the Far North, even if it was at times extreme. My first flight as a 13-year-old CAP cadet was in a bedraggled 85 hp not-so-Super Cub on skis out of Anchorage’s Merrill Field. Later, during my sophomore year at the University of Alaska in Fairbanks, I flew in a ski-equipped Cessna 180 on several wolf-counting missions in temps that often dropped to -30 degrees C.

I thought it was all a big adventure, and it never occurred to me that some people would stop flying their airplanes in winter, just because it was cold and there was snow on the ground. CAP flew all year round on search-and-rescue missions, mostly search and rarely rescue.

Depending upon where you live, winter can be an incredibly beautiful time to fly. Generally speaking, the sky is crisper, cleaner and colder, offering better visibility than in summer; the airplane performs better than during the warmer months because of the denser air, and strong seasonal winds sometimes can whisk your airplane to your destination in less time than usual.

Flying in winter does demand a few special precautions, however. You may need to review some basic pilot techniques and adopt a few new ones if you hope to realize the benefits of winter flying without some of the problems that go with it.

Keeping the airplane in a hangar is an obvious choice to avoid accumulation of ice and snow. Pilots who hangar their airplanes at home base often don’t worry about providing the same protection on a long trip. It’s perhaps ironic that conditions demanding a more thorough preflight often result in an abbreviated walkaround.

A lighted, heated hangar expedites the preflight, reduces the temptation to skip some items and streamlines the process of preparing the airplane to fly. Even an unheated hangar is better than leaving your bird sitting on the ramp overnight in frigid conditions.

A hangar will prevent frost or snow from accumulating on the wings and tail, and keep the airplane aerodynamically flyable. It will also preclude snow and ice from jamming or plugging pitot tubes, static ports and air inlets.

I ferried a Marchetti SF-260 from Santa Monica to Coventry, UK, several years back by the usual Goose Bay, Narsarsuaq, Reykjavik route. Sadly, there was no more room in the large group hangar at Reykjavik when I arrived, so the Marchetti had to sit out.

A lighted, heated hangar expedites the preflight, reduces the temptation to skip some items and streamlines the process of preparing the airplane to fly. Even an unheated hangar is better than leaving your bird sitting on the ramp overnight in frigid conditions.

The following morning, there was a thick layer of ice and snow on the airplane, effectively contaminating the airfoil. Since the deice truck was out of commission, the only viable solution was to use my credit cards to clear the wings and tail of frozen residue. I was careful to choose only those cards I knew I probably wouldn’t need on the remainder of the trip. Sure enough, I broke three cards in the 90-minute de-ice process. (For the record, my Chevron card was the strongest.)

Incidentally, don’t even consider using antifreeze to dissolve snow and ice on a painted, parked airplane. Antifreeze consists primarily of ethylene glycol, which will discolor paint. If your airplane is polished aluminum, there’s no paint to damage, but spilled antifreeze is bad for the environment, and it will poison any animal that consumes it.

One of the obvious benefits of a heated hangar, assuming you can find one on the road, is that you’ll save the money you would’ve paid for a preheat or a battery jump. This will make engine start as easy as it would be in Southern California in April.

A cold-soaked airplane, left outside overnight in temperatures below -5 degrees C, deserves both a preheat and battery jump. Some pilots argue that you really don’t need a preheat until the temperature drops to -15C, but I figure if I take the best possible care of my engine, it will take care of me.

I’m no A&P mechanic, but friends who are tell me that wear and tear on moving engine parts in a cold start just isn’t worth the money you’ll save skipping the preheat. Bearings, rings, crankshafts and other moving parts don’t last long in those conditions.

Preheat typically demands less than an hour while you file a flight plan, tend to customs, pay for fuel and services, and drink coffee. Thirty to 60 minutes of heat will usually bring the engine to a temperature that should allow a normal start.

Incidentally, try to resist the temptation to have an extra preheat hose fed through the pilot’s storm window to warm the cabin interior. Preheaters sometimes emit carbon monoxide, and that’s the last thing a pilot needs to have floating around inside an aircraft cockpit.

Once the engine is running, be cautious when taxiing on a snow/ice surface, especially in extreme conditions. When temperatures subceed about -30 degrees C overnight, tires may freeze a flat spot on the bottom. Add major power to taxi, and you may start chipping rubber on the corners of the flat spot when the tire begins to roll. Consider having the lineman dedicate a few extra minutes to each tire during preheat. Either way, it’s better to induce taxi with minimal throttle to barely get the wheels moving and provide the rubber a chance to expand slowly.

Breakaway thrust will probably be higher, and you may need more power to maintain proper steering in snow. If you’re taxiing in slush, keep the speed low to avoid splashing water onto the tires, brakes or gear mechanism. Keep in mind the brakes may be locked up initially if you applied the parking brake.

I once landed the first Piper Mirage (headed for Germany) at Goose Bay, Labrador at about -25 degrees C, only to discover the brakes were totally frozen. Fortunately, the runway was still covered with snow, and the Piper skidded to a stop without damage to the tires.

Similarly, taxiway lines may be obscured with snow or ice, and snow accumulations may cause other airplanes to be parked without the usual clearance from the taxiway centerline.

Regardless of your normal practice, leave the flaps full up during taxi to avoid blowing snow or freezing water into hinge tracks. Try to maintain a greater interval between you and any traffic ahead, since directional control may be marginal.

When you push the throttle up for takeoff behind a normally aspirated engine in extreme low temp conditions, at least consider the effect cold, compressed air may have on power. If you’re flying from near sea level at, say, -35 degrees C, the density altitude will be roughly minus 4,000 feet. That’s 4,000 feet below sea level. That means full throttle may deliver 33 inches of manifold pressure rather than the normal 29. Consider monitoring power to keep the mp below 30 inches.

Perhaps surprisingly, bird strike hazards can be higher in winter, especially on asphalt runways after sunset. Asphalt tends to absorb and retain heat during the day, a factor that can attract birds not smart enough to fly south for the winter. No matter what combination of lights you normally use for takeoffs, it’s best to turn on everything for a cold weather departure day or night. Once you’re safely off and up, you can power down the electrons to save the landing lights.

Another good idea on retractable-gear models is to leave the wheels down for an extra 30 seconds after liftoff to blow any water or accumulated snow off the undercarriage.

If you’re willing to fly relatively high, winter winds may provide a welcome push, especially if you’re flying east. True jet streams rarely reach much below FL200, but breezes at lower levels will usually compliment the winds up high. If the wind isn’t going your way, you can cut your losses by flying low. (Back in March 1994, Mooney loaned me a new Bravo, and I took advantage of spectacular tailwinds at FL250. I flew coast-to-coast from Los Angeles to Jacksonville in seven hours nine minutes at an average 300.2 mph, setting eight, world, city-to-city speed records in class C1C and unlimited in the process. That would have been highly unlikely in summer.)

Returning to Earth can be a little different in winter. Falling snow can render landing lights worse than useless during a night approach. You may actually find forward visibility better without them, as the high intensity light tends to scatter off the snowflakes and bounce back at you. Better to avoid fast approaches to slick runways and plant the airplane firmly on the asphalt rather than try to grease it.

Stay off the brakes if there’s any question of ice under the snow, and when you do apply them, do so gently and evenly. Never try to brake and turn at the same time, especially if you’re flying a tricycle with the steering wheel out front. An iced runway isn’t the place to practice short field procedures, especially if black ice conditions prevail.

Finally, don’t be afraid to say to hell with it and go someplace else if conditions aren’t to your liking. I’ve made enough dumb decisions on snowy runways to scare myself several times, fortunately without damage.

Winter flying, like aviating in any other season, has its peculiarities, but snow pilots are convinced the rewards are worth the effort. If you’re forewarned with a little knowledge of winter’s potential pitfalls, you can enjoy flying when the sun and the temperature are low.

Dealing With Inflight Icing

Icing terrifies many pilots. Under the worst circumstances, it can turn a beautiful, aerodynamically clean aircraft design into a flying brick with little possibility of control. Fortunately, icing hazards can be minimized with intelligent decision-making and strict adherence to the pilot’s universal rule, “Fly the airplane,” and the first corollary, “Don’t panic.”

You can sometimes negate some of that fear at major airports with ground deicing, though that’s an expensive treatment that’s not always available at smaller airports. If you do use ground anti-icing, try to avoid the fate of Air Florida’s Palm 90 at Washington National in 1982. The 737 was deiced, then delayed 50 minutes and again covered with ice during the wait. The crew departed anyway, the airplane stalled on takeoff and crashed into the Potomac River, clipping a bridge in the process. For that reason, try to get into the air as soon as possible after application, or you may need to repeat the process. The Air Florida crew needed to but didn’t.

A more expensive alternative is TKS, an inevitable abbreviation for Tecalimus, Kilchrist and Smyth, inventors of the system sometimes better known as the “weeping wing.” A good friend in Boston installed an aftermarket TKS system in his Mooney 201. Reed flew regular IFR on business trips around New England year ’round, and made it a habit to top the six-gallon TKS tank if there was any possibility of in-flight icing. He would simply select the anti-ice position—that dispenses about 1.5 gph—right after engine start. By the time he pushed the power up for takeoff, the wings would be coated, and he could fly relatively impervious to the danger of icing.

If he did encounter extreme conditions, he’d merely select the deice setting, which doubles the TKS flow rate, and that would practically guarantee no ice could adhere to the airplane, though available TKS time would be cut in half.

In delivering new and used aircraft across the North Atlantic since 1977, I’ve had TKS protection a few dozen times but rarely needed it if I was careful in selecting a cruise altitude. Traditional wisdom suggests icing is only a concern inside a temperature envelope of -20 degrees C to +2 degrees C. (Remember, OAT gauges are even less reliable than fuel gauges.) For that reason, most experienced pilots try to operate above the weather in winter, or to climb above it when they experience icing.

Ideally, you’ll need a clearance for that, but don’t wait too long. While ferrying an Aerostar 700 from Tousous le Noble, France to Cape Cod 20 years ago, I was held at 8,000 feet in rime ice over the North Sea for 20 minutes, despite three requests for higher. Finally, I advised the controller I was going to declare an emergency and ascend to my assigned FL160 if he didn’t immediately clear me to a higher altitude out of the ice. I got my clearance in 20 seconds.

The alternative to TKS is pneumatic boots on the wing and tail-leading edges, usually with alcohol deice systems for the props. Boots aren’t as effective as TKS, but if you understand their limitations (primarily, don’t activate them too early), they’ll break off ice in chunks and help keep you safe.

Keep in mind that most icing is confined to narrow, vertical bands of cold, moist air, often no more than 500 feet thick and 10 miles across. If you stumble into ice accidentally, you’ll probably fly out of it in a few minutes.

The normal rule in icing conditions is to climb first, while you still can. If you ascend, and things don’t improve, you can always descend. Descend first, and you may lose the option of climbing back up.

If there’s any good news, it may be that in-flight icing, whether it’s rime or clear, is generally rare and most often easy to avoid. The temperature limits favor semi-low altitudes where moisture content is higher, along with fall and spring when temperatures aren’t so cold as to completely preclude icing. Winter isn’t the most popular season for ice in flight.

The post Learning To Love (Or At Least Accept) Winter appeared first on Plane & Pilot Magazine.

]]>
Weather Avoidance Techniques https://www.planeandpilotmag.com/article/weather-avoidance-technique/ Tue, 12 Mar 2013 04:00:00 +0000 http://planepilotdev.wpengine.com/article/weather-avoidance-technique Staying Out of Trouble When it isn’t CAVU

The post Weather Avoidance Techniques appeared first on Plane & Pilot Magazine.

]]>

Some days, weather isn’t a problem: “ceiling and visibility unlimited” (CAVU). But, if you use an airplane for more than joyrides, there will be other days when getting from point A to point B will require flying over, under or through clouds. There can be bad things in those clouds: severe turbulence, icing and embedded thunderstorms. Today, a wide range of tools are available to help avoid those dangers, but each has limitations that must be understood in order to use them safely.

The most fundamental tool is available to all pilots: It’s the Mark I Eyeball. Most of us have two, and if you’re flying with a copilot or passengers, you may have spares available. It’s easy to use and has just one significant limitation: You have to look outside the airplane. Stay out of the clouds, and you’re almost guaranteed to avoid the worst effects of weather. This is a useful technique not only for visual flight rules (VFR) pilots, but also for instrument flight rules (IFR)—operating on top or between the layers, where you can see, is an ideal way to avoid nasty surprises.

Another tool is available to any pilot with a working radio: Flight Watch, which you can reach on VHF frequency 122.0 (when calling, give the nearest VOR: “Flight watch, N12345, Fortuna”). This will connect you with a flight service specialist who has access to current weather, including radar imagery.

Tell the specialist where you are (“12 miles east of Fortuna on the 260 radial, 6,000 feet”) and where you’re going, and you’ll get current conditions, Airmen’s Metereological Information (AIRMETs) and Significant Metereological Information (SIGMETS) and an outlook for your destination. On cross-country flights, I generally try to contact Flight Watch at least once on each leg—hourly on long legs. More than once, I’ve been alerted to problems in time to take action before a situation became serious.

For IFR pilots, an additional (and very important) tool is whatever ATC facility you’re in contact with—center, approach, departure or a local tower. Most ATC facilities have radar with some weather capabilities. They’re also in contact with other airplanes, which may be flying ahead of you on the same route.

If it’s getting rough, precipitation is increasing and you’re starting to wonder if continuing on your current route is really such a good idea, ask ATC if anyone’s ahead of you—and how it’s working for them—or what the weather looks like up ahead on the radar. If an airliner ahead is reporting a rough ride, or the last couple of airplanes missed the approach, or a suspected thunderstorm cell is moving across your route, it would be nice to know ahead of time!

Beyond the Mark I Eyeball and helpful folks on the radio, a number of devices are available that give the pilot a picture of the weather. Of these, the most common (and least expensive) is data link weather, provided either by satellite or a ground-based system, such as the FAA’s Flight Information Service (FIS), a component of Automatic Dependent Surveillance (ADS/B).



The timestamp in the upper left corner says this NEXRAD mosaic is only two minutes old—but it can be based on radar data more than ten minutes old.

Data link systems put a wide range of weather products on the multifunction display (MFD) in your panel, or a portable display. Depending on the particular system, this can include the Next Generation Radar (NEXRAD) radar mosaics, or Aviation Routine Weather Report (ETAR). It might also include the Terminal Area Forecast (TAF), in text or graphical form, AIRMET/SIGMET and some very useful non-weather data, such as temporary flight restrictions (TFRs).

But, there’s one big catch: Data link weather is never completely up to date. The service provider receives data from a variety of sources, puts it in a form appropriate to their system and transmits it (via satellite- or ground-based radio link).

The equipment on your airplane receives the signal, decodes it and displays it. The process takes a minimum of several seconds, and in many cases, the data can be tens of minutes old—something that should come as no surprise to any pilot who’s called up the Automatic Terminal Information Service (ATIS) for an airport just before the hourly update.

This limitation (referred to as latency) is a particular problem with NEXRAD radar imagery. Any data link system that displays NEXRAD will show you the time that the mosaic was created, which typically will be about five minutes old—but that’s misleading: The mosaic itself is based on data from a number of different radar systems, which have to be collated together. The actual data on which a NEXRAD mosaic is based can be as much as 20 minutes old, and there’s no way to tell that from the display.

Twenty minutes may not sound like a long time—but thunderstorm cells can build at rates of several thousand feet per minute, and cells in a line can advance by miles in that time. What looks like a safe route between a couple of small cells on a data link display may turn out to contain weather that no airplane can successfully fly through (for more on this topic—including details on two fatal crashes—see “When Using NEXRAD Can Be Dangerous” from the September issue of Plane & Pilot, and the links in this article’s sidebar.)

Because of latency, data link weather should never be used by itself for tactical decision making at short ranges. On the other hand, data link weather is an outstanding tool for strategic decision making at long ranges—if you’re flying a piston single or light twin without onboard real-time weather avoidance gear (and flight into known icing equipment) the best way to deal with a line of thunderstorms is to change your route early and fly around the entire area, giving yourself at least 20 miles of space on the upwind side.

To fly through a line, you need equipment that will show you what the weather’s like in real time: either onboard weather radar or lightning detectors. Each has limitations.

Onboard Radar
Before flying an airplane equipped with onboard radar into active weather, I urge you to take one of the available radar training courses, such as Sporty’s excellent Airborne Radar Training Course, or find a copy of Airborne Weather Radar: A User’s Guide by former Delta captain James C. Barr (it’s out of print, but available from online bookstores). On-board weather radar is complicated.


To use it effectively, you’ll need to understand antenna tilt, attenuation, how the manufacturer maps intensity to color, storm gradient and shape, range settings, gain, how to estimate radar tops and optional features like stabilization, turbulence detection, wind shear detection and sensitivity timing control. Failing to understand these features may result in misinterpreting the radar picture and flying into an area where you really don’t want to be!

One key thing to understand is that any radar, whether airborne or ground-based (including NEXRAD) detects only water. It can’t show clear air turbulence, and also can’t detect ice or dry hail. Tops of thunderstorms at very high altitudes may not be detectable as they consist almost entirely of ice.

Lightning Detection
As a practical matter, onboard radar is available only on multi-engine aircraft and the largest singles. It’s expensive, and radar performance improves with antenna size—the smallest available is generally around 10 inches in diameter. You’re not going to fit something like that on a typical piston single. An alternative technology that can be used on any airplane with an electrical system is lightning detection.

Lightning detectors are essentially sophisticated directional radio receivers that listen for the radio noise emitted by lightning strokes. They estimate distance based on the strength of the noise pulse. The result is a display that shows approximately where the lightning is relative to your aircraft. It has two advantages over radar-based systems: It works whenever there’s lightning, whether there’s water present or not, and it gives you a 360-degree view, showing activity to the sides and even behind you (onboard radar generally is limited to a narrow arc directly ahead).

Just as with data link and onboard radar, onboard lightning detection has limitations. The most significant is that only cells generating lightning are shown. Lightning detectors can’t warn you about heavy rain (or other precipitation) that isn’t generating lightning. This equipment also has a range limit, and there are other issues that vary from one type of detector to the next—you’ll want to study the manual before flying with one in heavy weather.

Practice And Plan
For all types of sensors, practice is essential whether you have radar, data link, lightning detection or some combination of different sensors, you should use them regularly—even on flights where you don’t expect to encounter significant weather. You don’t want to find yourself trying to figure out the fine points of your equipment while in the soup on a rough day!

Before departure on any flight where significant weather is expected, get a complete preflight briefing so that you know what to expect—and make sure whatever weather avoidance equipment you have is working: Preflight inspection of radomes/antennas and ground tests of each sensor are critical before takeoff.

If you’re launching into active weather, turn on whatever sensor(s) you have, but bear in mind that data link weather can be old—if you’re looking at cells, check the age of data and think through how they may have moved/developed.


For on-ship radar, turn the system on and tilt the antenna up to paint any calls on your departure path. If you have a lightning detector, turn it on before taxi. You may also want to advise ATC (ground or tower as appropriate) that you may need extra time on the runway to stabilize the radar or other sensor before rolling for takeoff.

There can be situations where takeoff is possible, but returning to the field won’t be (cells on the approach path). In that case, you may want to identify a takeoff alternate in case problems develop during climbout.

Deviations around active cells can be worked out with ATC, based on information from whatever sensor you have —but if you’re displaying on a multifunction display (MFD) and operating over anything other than flat terrain, you may want to periodically switch between the radar/data link/lightning display and terrain. You don’t want to deviate around a cell and fly into a mountain!

As soon as you think a deviation may be needed, advise ATC, so that they can identify any conflicting traffic and approve or deny your request early.

With on-ship radar or lightning detectors, it’s important to periodically switch to maximum range while working around local cells—otherwise, you might succeed in getting around a small problem and wind up face-to-face with a bigger one.

Thunderstorms are dynamic, and can build at rates of 5,000 FPM or higher. Thus, old data—from any sensor—can be worse than useless, tempting you to fly into areas that look clear, only to encounter a rapidly developing monster. If you’re using data link, the only way to avoid this is to give all active areas a wide berth. With onboard radar, periodically checking shorter ranges may help. Lightning detectors will show any cells developed enough to produce lightning, but may not show building cells that haven’t had time to develop significant electrical activity.

As you plan your approach, use whatever sensor you have to inspect weather not only at the destination airport, but also on your missed approach. If there’s a cell near the missed approach hold you’ll want to know about it and discuss what to do with ATC before beginning descent.

Understand the capabilities and limitations of whatever weather avoidance equipment you have, and you can fly safely when conditions are well below CAVU!

For More Information
NTSB Safety Alert on NEXRAD Mosaic Imagery
www.ntsb.gov/doclib/safetyalerts/SA_017.pdf

Aeronautical Information Manual 7.1 (meteorology)
www.faa.gov/air_traffic/publications/atpubs/aim/aim0701.html

Airborne Radar Training Course (CD and online versions
www.sportys.com/pilotshop/product/16954

Sirius XM Aviation Weather (data link)
www.xmwxweather.com/aviation/

Avidyne TWX670 Tactical Weather Detection System (lightning detector)
www.avidyne.com/products/twx670/index.asp

L-3 Avionics Systems Stormscope (lightning detector)
www.as.l-3com.com/products/stormscope

Insight Avionics Strike Finder (lightning detector)
www.insightavionics.com/strikefinder.htm

Honeywell Bendix-King ART-2000 (on-board radar)
www.bendixking.com/ProductDetail?ProdName=art-2000

The post Weather Avoidance Techniques appeared first on Plane & Pilot Magazine.

]]>
10 Tips For VFR Flying In Marginal Weather https://www.planeandpilotmag.com/article/10-tips-for-vfr-flying-in-marginal-weather/ Tue, 05 Feb 2013 04:00:00 +0000 http://planepilotdev.wpengine.com/article/10-tips-for-vfr-flying-in-marginal-weather A few simple suggestions to help keep VFR pilots safe in marginal weather

The post 10 Tips For VFR Flying In Marginal Weather appeared first on Plane & Pilot Magazine.

]]>

It was the classic example of baby- bird stupidity. I had flown myself into a box of clouds without the benefit of training, experience, or even a properly equipped airplane. I was a two-year private pilot at the time, flying above the mountains of Arizona on my way to Albuquerque, N.M. Although I knew exactly where I was, that only served to define where I’d probably crash.

My first airplane, a defenseless, little Globe Swift, wasn’t even close to being instrument flight rules-equipped (IFR). It had no artificial horizon, a venturi-powered directional gyro (DG) to tell me where my plane’s nose was pointed and a VOR (a short wave radio navigation system) that only worked on alternate Thursdays in March. My instrument proficiency consisted of about 1.5 hours under the hood, acquired two years ago, in preparation for the private pilot test.

The fact that I’m writing this suggests the ending was anticlimactic. With clouds coming down and the terrain coming up, I finally identified where I was along Route 66 and turned to follow it east as it climbed toward the overcast. Within a few minutes, I spotted the barely visible, unimproved short strip known as Dinosaur Caverns. I had seen it several times before. (Don’t bother to look it up. It’s been closed since 1975.)

I landed the little Swift on the sagebrush-strewn runway, tall weeds scraping the bottoms of the wings, and rolled out onto a small ramp that also served as a parking lot for a local gas station.

Although I had done practically everything wrong, somehow science and technology had triumphed over fear and superstition. There was also a major amount of blind, dumb luck.

Quite a few flight hours and a number of ratings later (including an instrument ticket), I’ve learned a few things, although not nearly enough. I’m determined never to make those same mistakes again. I’ll find some new ones.

One benefit of writing for the magazines is that I’ve been given the privilege of flying with 100 or more check pilots in conjunction with pilot reports. That’s given me a virtual cornucopia of ideas on all things aviation, especially the pitfalls of flying into IFR conditions without proper training or equipment.

The following are some of the suggestions made by these instructors, check and test pilots, bush pilots and miscellaneous aviation bums who very well may know more about flying than I could ever imagine. Most of these suggestions are more common sense than revelation, more homegrown flying philosophy than scientific fact. Take them for what they’re worth.



Unless you’re instrument rated, consider staying on the ground when conditions are marginal.

1 Don’t judge weather for a cross-country trip by simply looking out the window, seeing clear skies and assuming it will be just as good at your destination. It’s ironic that pilots, supposedly trained to make dispassionate judgments on weather, sometimes make illogical decisions.

Just because it’s ceiling and visibility unlimited (CAVU) where you are, don’t be seduced into assuming that’s how it is at your destination. Such conclusions become especially relevant when the distance stretches to 700 miles or more and you may be flying adjacent to several weather systems.

Several years back, I was delivering a new Partenavia P-68 Observer from Naples, Italy, to Santa Paula, Calif., an airplane notorious for its intolerance of carburetor ice. Accordingly, I was forced to luxuriate under clear skies at Reykjavik, Iceland, for six days while winter storms dropped a ton of snow on Greenland and Labrador, Canada. I imagine I could have flown out a few hundred miles to “take a look,” but it would have been a waste of very expensive avgas.

2 Be especially wary of any forecast that suggests marginal weather at your proposed destination will remain the same when you plan to arrive several hours later. Remember that “about the same” isn’t far from “deteriorating to!” If the forecast suggests conditions may be “improving to!” by your expected time of arrival (ETA), you may stand a better chance of completing the flight without problems.

One question I’ve been trained to ask the briefer on every trip is, “Where is it good?” If the weather seems to fall on the positive side of my decision process, and I’m inclined to give it a try, I’ll often ask the briefer for suggestions of alternates along my route or possibly a different route altogether that might avoid the nasty weather in the first place. That might give me more options, should I decide to park the airplane and wait for better conditions.

3 Think at least three times about flying VFR at night. A few years back, the FAA considered a requirement for extra hours of simulated instrument training for those VFR pilots flying at night. Ten hours was a popular number. Apparently, no one felt there needed to be a special VFR/night rating, but everyone agrees night VFR is more challenging than day operation. There’s often no visible horizon at night, you generally can’t see clouds, and if you’re operating over remote areas where lights are sparse, you may not be able to differentiate ground details at all. Unless there’s moonlight, night operation can simulate a black hole, and that’s no place for a VFR pilot. On top of that, there are certain aircraft sounds that can be heard only at night. If that spooks you, perhaps you should stick to flying in daylight hours.

4 Unless you’re IFR rated and totally up to speed, don’t even consider flying in weather reported at 1,000 and three. Yes, technically, you’d be legal, as long as you’re in the pattern, but it’s hard to imagine what you could accomplish in such marginal conditions.


Assuming you’re flying above what the FAA calls a “congested area,” you’ll need 1,000 feet above ground and 500 feet below the clouds. That’s 1,500 feet above ground level (AGL), and even that isn’t enough. Generally speaking, if the forecast along your route or at your destination doesn’t suggest the ceiling will be at least 2,500 feet and the visibility five miles or better, be wary.

I once delivered a purely visual flight rule-equipped (VFR) Pitts S2C from Dallas, Texas to Long Beach, Calif., over three days with nine stops for the 1,050 nm route, deviating three times for weather. It wasn’t pretty, but it worked, and both the airplane and I arrived in one piece. It did cost the client a few extra bucks, but because I was willing to reroute and delay as necessary, I got the trip done safely.

Just because it’s ceiling and visibility unlimited where you are, don’t be seduced into assuming that’s how it is at your destination.

5 Temper your judgment even further if you’re flying in mountainous terrain. I’ve lost so many friends to combinations of clouds and mountains in the last 40 years that I’ve become more than a little paranoid about mountain flying with clouds nearby. A few years back, two highly experienced, instrument flight rule-licensed IFR pilots, one of them a good friend, flew a new Caravan straight into the side of a mountain in California’s Banning Pass, apparently trying to stay out of the clouds and avoid icy conditions.

Every pilot has the right to decide what’s an acceptable risk (if he’s flying solo), but I don’t even consider launching into the mountains even in slightly questionable conditions, regardless of whether there’s a turbo out in front or not. Mountain passes aren’t necessarily a better alternative either because clouds can close in behind and leave you with no way out.

6 If you’ve made a decision to launch toward atmospherics that you have the slightest question about, get an update long before you fly anywhere near the bad weather to minimize the chance of an inadvertent bout with something you can’t handle. If you’re not IFR rated and never fly into IFR conditions in the first place, you’ll never have to worry about escaping from them.

7 One comment you hear repeatedly from many pilots concerned about flying in inclement weather is that an instrument rating only encourages aviators to operate in conditions they can’t handle. Such a philosophy is fine, as long as you subscribe religiously to pure VFR and forgo use of your airplane when clouds are about.

Regardless of whether you have plans to earn an instrument rating, make it a habit to spend a little hood time now and then, to refresh your proficiency, just in case. If your total instrument training is the minimal emergency exposure for the private ticket, you may stand little chance of surviving a real IFR encounter.


8Every pilot without an instrument rating fears winding up on top of an overcast, especially if you’re doing everything right and the insidious clouds sneak in beneath you.

This is a special risk over swamps or coastal areas where water can contribute to instant ground fog. Too many times, pilots insist on watching the sky rather than the ground and barely notice when clouds creep in insidiously and blot out the ground. Keep an eye on the bottom quadrant and be certain to maintain ground contact.

9 If everything goes down the tubes and you do accidentally blunder into IFR conditions, the quickest route to safety is usually a 180-degree turn. Usually. No need to wrap the airplane into a steep bank and risk vertigo. Make the turn standard rate, and in a minute, you’ll probably be headed out of trouble. After all, you got into this mess in VFR conditions, so a reverse track is often the safest way out.

Too many times, pilots who accidentally intrude on instrument metereological conditions (IMC) delude themselves that they can bluff it out by climbing through to on top or letting down a little below the clouds. Climbing through may solve the current problem but leave you with the obvious question of getting back down through the overcast. Letting down is nearly always a bad choice unless you know exactly where you are and what’s below. After all, down is where the ground lives.

10 Finally, should you wind up in the clouds, alone, scared and without a clue, call for help as soon as possible. Assuming you can maintain control and keep the airplane level, don’t blindly try to find your way out without help. Controllers aren’t magicians, but if you have some vague idea where you are, they may be able to identify you on radar and provide vectors to clear air and low terrain. Equally important, they can warn legitimate IFR traffic to stay out of your way.

Don’t assume just because you made some dumb mistakes that you’ll have automatically earned a Federal Aviation Regulations (FAR) violation. Most controllers are reasonable people who accept the fact that everyone makes mistakes, and if you’re honest about the situation and they’re convinced your poor judgment calls were exactly that and not deliberate recklessness, they’ll probably let it pass. Even if they don’t, a violation is better than the worst alternative.

The obvious solution to weather accidents is an instrument rating, and even that won’t solve all your weather problems. For those who had rather not spend the time and money to learn instrument flying, fine. There’s no disgrace in flying only when the skies are clement. Just don’t make the mistake of trying to mix IFR and VFR.

You may fool some of the weather gods some of the time, but you won’t fool all of them all of the time.

The post 10 Tips For VFR Flying In Marginal Weather appeared first on Plane & Pilot Magazine.

]]>
Combating Crosswinds https://www.planeandpilotmag.com/article/combating-crosswinds/ Tue, 06 Mar 2012 04:00:00 +0000 http://planepilotdev.wpengine.com/article/combating-crosswinds Ten Things The Textbooks Don’t Tell You

The post Combating Crosswinds appeared first on Plane & Pilot Magazine.

]]>

Here’s an aviation fact that you can take to the bank knowing it’s almost always true: Neither the wind given to you by the tower, nor that shown on a mid-field wind sock, is likely to be what you actually experience when landing. There are lots of reasons for that, and understanding those reasons will cut down the number of surprises and the amount of drama in your crosswind landings.

1 A Steady Wind Isn’t Even Remotely Steady If wind had color, so we could actually see how it’s shaped, we’d see that like a river, it’s interrupted by swirls and eddies. It’s constantly changing, flowing smooth for a few seconds, then changing direction and strength, only to go smooth again. So when we’re in the approach and, even more important, when we’re in flare and floating down the runway, we’re experiencing an ever-changing combination of wind-induced bumps and wiggles. Every several hundred feet down the runway promises a new experience.

2 Wind Is Made Up Of Layers When thinking about fighting wind, we often think of nothing but crosswinds, which is a two-dimensional, left-right way of thinking. However, wind, like airplanes, is three-dimensional, and is usually composed of several layers. It’s quite common to turn final and find that our glideslope has changed. At first we looked high, then for no apparent reason, we’re low (the numbers are moving up the windshield), and a few more ponies are needed to make the runway. Or, just the reverse happens. This is because the wind at pattern altitude is much less affected by either the friction wind develops with the ground, or the topography and other ground-bound obstacles that the wind has to work around. Also, thermal effects are generally more intense closer to the surface. So, don’t be surprised to find that the wind on the ground is grossly different than what was experienced on downwind.

3 Topography Can Really Affect Wind On Short Final Not all runways are in the middle of a mile-square, billiard-like wheat field. In fact, in many parts of the country, runways are perched on, carved out of, clinging to or stuffed into real estate that’s everything but flat. And that being the case, the wind’s path can be pretty screwy because of what it’s curling over and around.

A curl is developed when the wind has to avoid an object and, in effect, wraps itself around the corner, e.g. a runway with a noticeable drop-off at the end. When the wind comes whistling down the runway and the pavement is suddenly no longer there, the wind curls downward. If we’re on short final, the curl will try to take us with it.

The reverse of the downward curl is when we’re coming over an obstacle at the threshold directly into the teeth of the wind that’s trying to climb over that obstacle. It will pick us up, then just as quickly, decide it no longer wants us, and where we were fighting an upper, we’re suddenly falling out of the air.
Another variation of curl occurs when we’re landing in a crosswind—a time when we really don’t need any more complications—on a runway with a ridge or tree line alongside it. This time, the curl is coming over the trees, and it may try to slam you into the ground. It may try to pick you up. It may just pound you senseless, then smooth out, as if saying, “Okay, I’ve screwed with you enough.”

4 Wind Socks At The Beginning Of The Runway Are Best We always try to land at the beginning of the runway (if we’re doing our jobs anyway). Unfortunately, most airports think we land in the middle of the runway because that’s where they put the wind sock. Socks can’t be that expensive! Every runway should have a sock at each end, so we actually know what the wind is doing where we’ll be touching down. The further down the runway a wind sock is located, the closer to fantasy it becomes.


Some specifics: You’re landing on a 5,000-foot runway and the only sock is mid-runway. That’s nearly a half-mile from where we’ll land. We can fit a lot of wind changes into a half-mile. Especially if the wind is crossed, and there’s topography or buildings out there adding their little bit of entertainment to the situation.

If wind had color, so we could actually see how it’s shaped, we’d see that like a river, it’s interrupted by swirls and eddies.

5 Control Towers Often Get Wind Information From Off-Airport Sources It’s common for the source of wind information broadcast on ATIS, especially at larger airports, to be off of the airport property. So, this time it’s an 8,000-foot runway, and the wind information is coming from a tower that’s a quarter-mile away: Now, we’re landing nearly two miles from the wind information source. Yet another reason why we search out secondary wind information sources, such as wind socks, flags, blowing dirt, trees, etc.

6 Beware The Down-To-The-Surface Wind One of the more unusual winds is the one that ignores boundary layer effects. As a normal rule, the wind right on the surface of the runway is zero, and it builds up to the reported velocity at about 15 feet. However, there are some “rogue” winds that hold much of their velocity right down to the pavement. This is just the reverse of what normally happens. In this case, the airplane is slowing down and losing energy, while the wind is doing neither. Because the wind hasn’t decreased, its effect on the airplane is actually stronger. These can be really, really nasty winds because they sneak up on you, and you don’t realize until the last second that it takes much more control movement than normal to beat them into submission.


The topography of a runway can change the path of the wind, sometimes quite dramatically. A wind curl is developed when the wind has to avoid an object and in doing so, it wraps itself around the corner.

7 Violent Changes In Direction Can Steal Lift One of the most dangerous winds doesn’t even have to be a big one. It can be less than 15 knots and still cause plenty of heartburn. It will have two characteristics that fortunately aren’t combined that often, but the combination occurs often enough that it’s worth discussing.

The first characteristic is sharp-edged gusts that change direction in a heartbeat. Making it much worse, the sharp peaks of the harder-than-normal gusts are off the main wind heading. As the gust dies, the wind returns to the main heading. So, every time it punches you, it’s from a different direction at a higher velocity.

The second characteristic is that the basic direction of the wind is 90 degrees to the runway. When a 90-degree wind is combined with sharp gusts that snap back and forth during flair, the possibility of gusts instantaneously changing from in front of the wing tip to behind the wing tip becomes critical. When wind snaps from in front to behind the wingtip, the airplane doesn’t have time to react to the sudden change in airspeed because of inertia. For a moment, it loses so much lift that it becomes dynamic, as opposed to aerodynamic. At that point, it’s an ingot, not an airplane. This is when burying the throttle in the panel is called for. Not partial power. Full power, because nanoseconds count.

8 Wind And Turbulence Interact One of the more “fun” wind conditions (read that as teeth-shattering because of the violent ups and downs) is seen in the West more than any other place. This is when a hard-edged, gusty wind is combined with low-level turbulence caused by high ground temperatures and/or topographical effects. The resulting ride can be violent and unpredictable. In these conditions, our primary job is to determine exactly what attitude is required to counter the crosswind, and how to “firmly” maintain it without over doing the corrections. When fighting turbulence in gusty crosswinds, there’s a tendency to accidentally bring the ailerons back “past center.” By this, we mean that we’re supposed to be holding a wing down, but without realizing it, we overcontrol in turbulence and actually put aileron in the other direction. This momentarily picks up the down wing.


When fighting those kinds of conditions,we have to visualize the exact wing attitude we need for the crosswind correction, then be very firm with the controls, making smooth, but quick jabs to maintain that attitude and keep the nose right in front of us. We don’t want the wind to fly the airplane. That’s our job.

9 Ground-Level Venturi Effects “Whoa! What was that?” We’ve all felt it on either landing or takeoff. Everything will be going just great, when we’re hit with something from the side that feels as if we’ve gone through a jet-engine blast. Usually, it’s a narrow steam of wind caused by a crosswind flowing between two obstacles (buildings, trees, etc.). The venturi formed by those obstacles causes the wind to narrow and accelerate, creating the jet-blast effect. These are quite often formed only during specific wind conditions, e.g. for a given airport, it may be a wind from 030 that happens to line up with a venturi-like gap between a building and a billboard, or something similar.

10 The “Character” Of The Wind Is As Important As Direction Or Velocity It’s not unusual to find ourselves working our butts off in a little 12-knot wind when we know that we’ve handled 25-knot winds with far less work. All winds aren’t created equal, and it’s the character of the wind that makes up much of the difference. Is it quirky with some sort of weirdly changing direction, or it’s three knots, gusting to 12 knots and the changes are instantaneous, so you often get left hanging? The personality types of wind are almost limitless. The good news is that by studying the wind sock and the local environment, you can get a rough idea of the general character of the wind you’ll be fighting.


The winds you actually experience when landing may be different from those that were reported. Factors such as topography and venturi effects will all come into play.

If the sock goes from practically limp to fairly straight, we know we’re going to have a bumpy, unpredictable ride. If its “stiffness” remains fairly constant, but it’s whipping back and forth, we’re going to be working hard to maintain an attitude. If it has very little wind in it, but it periodically circles the mast in a lazy fashion, there’s a chance we may be landing with a slight tailwind. If it’s at 90 degrees to the runway, with regular excursions to 120 degrees (30 degrees behind the wing) and is gusting 15 to 25 knots, it might be a good day to visit the airport restaurant and wait until Mother Nature makes up her mind.

Pay special attention to everything around the airport and the runway that can indicate wind. This includes flags, hanging decorations on businesses, tethered balloons, etc. If you see a flag that’s off-airport that indicates a wind that’s quite a bit different than the sock, you know something exciting is happening in between. If the grass next to the runway is laying flat, you know it’s one of those “nasties” that retains its velocity right down to the surface. Watch out!

Summary
A runway is nothing more than a long string of micro climates, little bubbles of wind, temperature, humidity, etc. that change as we travel down the runway from bubble to bubble. Knowing that arms us with one very important fact: Regardless of what the tower or wind sock tells us, we should be prepared to deal with whatever we’re seeing around the airplane at any given moment. The old adage, “Fly the airplane,” applies here. If something is making it do something you don’t want it to do, we don’t really care what’s causing it. We just do our pilot thing and put the airplane where we want it.

Incidentally, there’s no takeoff that absolutely has to be made. If the wind is too far out of your comfort zone, don’t fly. And we should never allow a fuel situation to develop that forces us into making a landing in a wind condition we think is over our heads. Always be prepared to go looking for an alternate airport with friendlier winds.

The post Combating Crosswinds appeared first on Plane & Pilot Magazine.

]]>