Structural Fire Design: Wood

3y ago
49 Views
5 Downloads
666.58 KB
21 Pages
Last View : 20d ago
Last Download : 3m ago
Upload by : Pierre Damon
Transcription

United StatesDepartment ofAgricultureForest ServiceForestProductsLaboratoryResearchPaperFPL 450StructuralFire Design:WoodE.L. Schaffer

AbstractAnalytical procedures to predict the fire endurance ofstructural wood members have been developed worldwide.This research is reviewed for capability to predict the resultsof tests in North America and what considerations arenecessary to apply the information here. Critical researchneeds suggested include: (1) Investigation of load levels usedin reported tests, and parameters in analyses, for applicationto North American practice; (2) the effect of lumber grade onwood property response at elevated temperature; and (3)further effort in reliability-based design procedures so that thesafety of fire-exposed members and assemblies may bedetermined.Keywords: Structural design, structural members,timber/structural, wood, wood laminates, fire resistance, fireprotection, structural analysis, connections, joints,beams/structural, columns/structural, structural adhesives,softwoods, hardwoods, heat resistance, high temperaturetests, thermal degradation, mechanical properties.October 1984Schaffer, E. L. Structural fire design: Wood. Res. Pap. FPL 450. Madison,WI: U.S. Department of Agriculture, Forest Service, Forest ProductsLaboratory; 1984. 16 p.A limited number of free copies of this publication are available to the publicfrom the Forest Products Laboratory, P.O. Box 5130, Madison, WI 53705.Laboratory publications are sent to over 1,000 libraries in the United States andelsewhere.The Laboratory is maintained in cooperation with the University of Wisconsin.

NomenclatureContentsPageIntroduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1Properties of Wood . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3Charring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3Temperature and Gradients . . . . . . . . . . . . . . . . . . . . . . . . . 4Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5Modulus of Elasticity (Parallel to Grain) . . . . . . . . . . . . 5Tensile and Compressive Strength (Parallel toGrain) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5DurationofLoad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6Other Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6Deformation(Time-Dependent) . . . . . . . . . . . . . . . . . . . . . 6FlreEndurancePrediction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8Beams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8areabeam breadthbeam depthcharring ratecorrection factor for column load and slendernessdistance into section from section surfacedistance into wood from char-wood interfacefunctionalsheat fluxloadload-bearing area section modulusmodulus of elasticitymodulus of rupturemoment of applied loadradius of gyrationratio of initially applied load to initial criticalbuckling loadslenderness ratiospan or column lengthstrainstresstemperaturetemperature reduction factortemperature-time shift factorthermal conductivitythermal diffusivitythermal expansion coefficienttimevolumeColumns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .12Decking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .13Tension Members . . . . . . . . . . . . . . . . . . . . . . . . . . . . .14CombinedLoadMembers . . . . . . . . . . . . . . . . . . . . . .14Research Needs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .15References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .15i

Mathematical Symbolsxilambdascript -wood interfacecreepcriticaleffectiveheatinitial or reference conditionresidual section and strength propertysecondary conditionssurfacetension

StructuralFire Design:WoodE. L. SchafferForest Products Laboratory, Madison, Wis.IntroductionHeavy timber construction was recognized as having firesafeattributes by the Factory Mutual insurance companies in theearly 1800’s (8).1 Massive heavy timber construction, whichminimizes concealed spaces for hidden fire spread and allowsminimal combustible surface area, withstood severe fireswithout structural failure. With the inception of the fireendurance rating system employing American Society forTesting and MateriaIs (ASTM) E 119 (4) fire exposure tests,heavy timber type construction, of specified minimumdimensions, was considered equivalent to or better than othertypes of construction having a 1-hour fire endurance. Thisappeared to be a “grandfather clause” for acceptance of aproven system. With the key issue being the differencebetween “real” fire and “simulated” fire performance, thisallowance was a rational decision.Fire endurance is defined (3) as a measure of the elapsed timeduring which a material or assembly continues to exhibit fireresistance under specified conditions of test and performance.AS applied to structural elements of buildings in NorthAmerica, it is measured by the methods and to the criteria ofASTM Standard E 119 (4). The structural members orassemblies are subjected to a standard fire exposure andevaluated for their continued load-carrying ability oreffectiveness to act as a heat transmission barrier. Singlestructural members are only evaluated for their load-carryingability. The standard specifies that the applied load be themaximum superimposed load allowed by design undernationally recognized structural design criteria. This loadingcondition is termed “full design load” and would bedetermined for timber constructions in the United Statesemploying the National Design Specification (39) and theTimber Construction Manual (1). The E 119 standard alsoallows test under less than “full design load” if suchrestricted load conditions are reported.The minimum nominal dimensions required for timber to beaccepted for classification as “Heavy Timber” are given intable 1. Though members were once exclusively sawn fromlarge-diameter logs, such sizes are now also available in gluedlaminated lumber (glulam) sections having equivalent fireperformance. Fire endurance tests of heavy timber members1Italicized numbers in parentheses refer to references cited atend of report.using the ASTM E 119 standard fire exposure indicate somesizes may not meet the performance requirements of thestandard for a l-hour rating.This paper will attempt to summarize the data base related tothe deterministic prediction and measurement of the fireendurance of heavy timber members. For testing the accuracyof analytical models, the characteristic loading conditions andactual fire endurance times of members are needed. This isdone for each member type.[ Note. United States-Canadian data cannot be directlycompared with Asian-European fire endurance data formembers. Though their fire exposure severities (timetemperature curves) are similar to United States-Canadianpractice (fig. 1), Asian-European countries compute allowabledesign stresses for the wood and members in markedlydiffering ways (18). Hence, for comparison, the AsianEuropean results must be translated to the United StatesCanadian basis. The mean strengths of dry clear wood, arereduced to design stress levels, fa, by applying a reductionfactor for variability, φ ; general adjustment factor, FS, thatincludes duration of load application effects; grade factor,GF; and cross-section size, CF:(1)The reduction factor attempts to correct a population toanticipated use of a weak member. In a statistically normalpopulation, these are usually 5 percent and 1 percentexclusion limits of strength (i.e., 95 pct and 99 pctrespectively of the wood used is expected to be stronger thanthis level). These factors are shown in table 2 (18) for severalvariability levels. The reduction factor for visually gradedlumber in the United States is about 0.474 and, for propercomparison, practices in other countries need to be calibratedto this level. The same care must be used in the generaladjustment factor, grade factor, and size factor terms,United States-Canadian FS levels for softwoods are stress typedependent as shown in table 3 (2). ]

Properties of WoodTo generate analytical models for estimating fire endurancethat are not simply empirical, the models need to includeparameters for the charring of the wood, compensation forwood strength or reformational characteristics at elevatedtemperature, and changing moisture content. Considerableprogress has been made in defining these effects in recentyears.Table 4.—Charring rate of vertically fire-exposed sections of coastDouglas-fir, southern pine, and white oak (52)CharringThe progressive conversion of the fire-exposed surfaces toever-deepening char occurs at definable rates. Because of thenegligible strength and fissured nature of this char, onlyuncharred wood is assumed to contribute to load-carryingcapability. The interface between charred and noncharredwood is the demarcation plane between black and brownmaterial. Because the temperature gradient through this areais steep, the demarcation is practically characterized by atemperature of 288 C (550 F).It is relatively well established that the rate of conversion tochar decreases with increasing moisture content and density ofthe wood used (45). Charring rate is also affected by thepermeability of the wood to gaseous or vapor flow. Charringnormal to the grain of wood is one-half that parallel to thegrain (19,22,54). As long as the residual section is large withrespect to the depth of char development, the rate isunaffected by the dimension of the section exposed.The charring rate, v, for vertically exposed surfaces of coastDouglas-fir and southern pine species (commonly used inglulam beams, columns, and decking) and white oak underASTM E 119 fire exposure (fig. 1) is given in table 4. Othercountries cite charring rates comparable to these for species ofsimilar densities. However, German experiments (29) haveshown that the bottoms of loaded beams experience a highercharring rate (0.043 in./min) during exposures of up to50 minutes. Evidently the increased charring is a result of theeffect of beam deflection to reduce insulative capacity of thechar layer. That is, the char layer develops wider fissuresthan in the nonloaded case.Charring rates have been both measured for various speciesand employed in design by various countries. In general,softwood rates range from 0.024 to 0.033 inch per minute(in./min) and are inversely proportional to density. Basedupon these results, a charring rate for all softwoods wouldconservatively be 0.031 in./min under fire exposure.Hardwood charring rates are less than 0.021 in./min.The charring rates cited apply to cases where members areeither large enough in cross section or durations of fireexposure short enough to minimize heat storage within theuncharred residual volume. A qualitative measure of theonset of heat storage is given by the time at whichtemperature at the center of a fire-exposed section begins torise significantly above that initially. A 2- by 4-inch section,for example, could tolerate only a few minutes of fireexposure on four sides, as compared to an 8 by 10, before asignificant heat storage effect develops. Such storage of heatwill increase the charring rate because less energy is requiredto raise the material temperature and more can be used inpyrolysis. For a given wood species, the energy stored withtime can be rigorously defined as a function of wood densityand specific heat capacity, member volume, surface areaexposed, and temperature difference between exterior andinterior. If all other variables are constant, one may expectthe time, t, until heat storage develops significantly to be onlya function of the member surface area exposed to fire, A S,and member volume, V:For a long beam or column, this can be expressed as afunction of initial fire-exposed perimeter and cross-sectionarea, A. For a three-sided fire exposure of a beam ofbreadth, b, and depth, d, the time is:(2)The relationship of charring rate to this effect has not beenqualified.3

Once the center of a section begins to increase in temperature,heat is being stored. In this case, too, no analytical solutionsare available to describe the temperature gradient change withtime.Temperature and Moisture GradientsThe temperature gradients generatedwood section are very steep becausediffusivity coefficient, α q, of wood.temperature range of 280 to 320 Cwithin a fire-exposedof the low thermalChar develops in the(536 to 608 F); 288 C(550 F) has been found to be a convenient temperature levelto locate the char-pyrolyzing wood interface through the useof embedded thermocouples. The steep temperature gradient(heat flux) generates movement of moisture within thesection. Description of the temperature and moisturegradients within fire-exposed wood sections has receivedconsiderable research attention in recent years. Suchdescription is intended to provide the basis for adjustingstandard mechanical properties for elevated temperature andmoisture content in fire-exposed load-bearing members.Providing an analysis that predicts either, or both, thetemperature gradient and moisture gradient within suchsections has not been attained to date (53). Though a finiteelement analysis does predict the temperature gradient quitewell in ovendry (0 pct moisture content (MC)) wood, theresults with moisture present do not. Approximations of thetemperature gradient at early and later stages of fire exposurehave been found useful. For fire exposure with little chardevelopment (up to 5 min), Carslaw and Jaeger (11) provideestimates for constant heat flux, qO:(3)where:(4)The heat flux,exposure.is about 3 watts/cm2 for a standard fireA second equation has been used (47) to describe practicallythe temperature distribution in the uncharred wood below thechar-wood interface at a distance, once a quasi-steady-statecharring rate, v has been reached. (This occurs about 15 to20 min after initiation of fire exposure.) The equation is:(5)The temperature distribution for times between 5 and15 minutes would require interpolation, as no satisfactorysolution is available.Kanury (25) provides estimates for the temperaturedistribution in solid panels exposed to fire on one side.Improved predictions of temperature and pyrolysis of woodare being sought (e.g., Kansa et al. (24)).The moisture distribution has been measured in sectionsduring and after fire exposure (13,47,53). One notes that themoisture decreases from a peak to zero in a 0.59-inch (1.5-cm)zone in the wood below the char-wood interface. Research(47, 53) has shown that a peak occurs at about 100 C and isabout 1.26 to 2.0 times greater than the initial MC. Thelocation of the peak is well correlated (R 0.98) to thelocation of the char-wood interface. Typical moisture andtemperature gradient curves are shown (fig. 2) for a southernpine section of mean dry specific gravity of 0.52 and initialMC of 10.0 percent.

StrengthThis section focuses on how various defect-free woodstrengths (tensile, compressive, bending, and shear) and themodulus of elasticity (E) are influenced by a change intemperature and MC. (Considerable recent research indicatesthat temperature and moisture change response of defect-freewood differs significantly from that of lumber and timberscontaining knots, checks, and slope-of-grain defects.Unfortunately there is yet no way to compensate directly forthe effect temperature and moisture have on defect-containinglumber. As a result, corrections for temperature andmoisture in structural lumber and timbers must be based upondefect-free response estimates.)Modulus of Elasticity (parallel to grain)The E of dry (0 pct MC) wood decreases linearly withincreasing temperature to about 200 C (fig. 3). Above200 C, there is some evidence it decreases nonlinearly. Forwood at 12 percent MC, a common in-use level, a small lineardecrease is observed to about 180 C, and decreases rapidlyabove this level (fig. 3).Tensile and CompressiveStrength (parallel to grain)The tensile strength parallel to grain exhibits a small lineardecrease to about 200 C; above 200 C the effect becomesgreater (fig. 4).Parallel-to-grain compressive strength of dry wood (0 pct MC)linearly decreases more rapidly with temperature than tensilestrength (fig. 5). Limited data for wood at 12 percent MCand temperatures to 70 C show an even greater decrease.5

Duration of LoadWood can carry substantially greater maximum loads forshort durations than for long durations. As a result theworking stresses are compensated for expected periods of loadapplication. The allowable stresses given in the NationalDesign Specification (39) have been adjusted to reflect theeffect of 10 continuous or accumulative years of full designload application and is termed normal duration of load. Theratio of other working stress levels to the normal allowablestress levels is shown in figure 6 (39). Note that for a periodof load application of full design load for 1 hour, theallowable normal stresses may be increased 47 percent. Theduration of load adjustment does not apply to moduli ofelasticity or rigidity.Other PropertiesFor detailed information on such other mechanical propertiesas shear strength and tensile strength (normal-to-grain), thereader is directed to a comprehensive survey produced byC. C. Gerhards (17).SummaryA rise in temperature decreases all mechanical properties andthe decrease becomes greater with increasing wood moisturecontent.The parallel-to-the-grain strength and stiffness responses may,at this point, be combined with temperature and MC gradientinformation for large fire-exposed sections. This is illustratedin figure 7 for parallel-to-grain E, and compressive and tensilestrength as a function of distance into the wood below thechar layer. The results apply to a cross section large enoughto minimize temperature rise at the center of the section andafter 20 minutes of fire exposure to allow a quasi-steadymoisture and temperature gradient to develop. These factorscan be applied to adjust the modulus of elasticity andexpected tensile-compressive strength for estimating rupturelevels under fire exposure. Care should be used in applyingany duration of load factor in accomplishing this. Toprecisely predict the true stress state, or predict failure, acomplete analysis including time-dependent stress-straincompatibility is required.Deformation (Time-Dependent)The parallel-to-grain time-dependent deformation (creep) ofwood is important to fire-exposed structural members.Though long-duration creep has been examined attemperatures of 25 C and several moisture contents, nosimilar long-term creep information is available at highertemperature with varying MC. Increasing the exposuretemperature results in increasing the rate of creep deformation(5,26,29,43,46). As MC is increased as well, the creep rate isincreased proportionately (6). Hence, hot moist conditionsare conducive to high creep deflection.6

whereThe effect of elevated temperature on creep response isreflected in the shift factor, aT. Creep increases dramaticallywith increasing temperature as shown by the response ofreciprocal shift factor with temperature shown in figure 8.Creep is magnified tenfold at 125 C and fiftyfold at 250 Ccompared to 25 C.Levels of creep are small at room temperature, but increasewith both temperature and MC (6) (fig. 9).Total creep strain behavior, can be prescribed as afunction of temperature, T, by a single exponential function(5):(6)where t is time in minutes, and T, the temperature in C.Such a form has been employed to predict th

wood property response at elevated temperature; and (3) further effort in reliability-based design procedures so that the safety of fire-exposed members and assemblies may be determined. Keywords: Structural design, structural members, timber/structural, wood, wood laminates, fire resistance, fire protection, structural analysis, connections .

Related Documents:

FIRE TOPPER Fire Bowl User Manual Home » FIRE TOPPER » FIRE TOPPER Fire Bowl User Manual Contents [ hide 1 FIRE TOPPER Fire Bowl 2 Setting Up Your Fire Topper Fire Bowl 2.1 Set-Up 3 Placement and Location 3.1 Liquid Propane Tank 4 Using your Fire Topper Fire Bowl - For your safety, read before lighting. 5 Cleaning, Maintenance, Storage 6 .

3 Key Advantages of Wood-based Fuels 1.1 Wood energy is widely used and renewable 1.2 Sustainable wood production safeguards forest functions 1.3 Wood energy is available locally 1.4 Wood energy provides employment and income 1.5 Wood energy supports domestic economies 1.6 Wood energy is modern and leads to innovation 1.7 Wood energy can make a country independent of energy imports

wood (see Section "Role of wood moisture content on corrosion"). 3. Wood preservatives Wood preservatives are chemicals that are injected into the wood to help the wood resist attack by decay fungi, mold, and/or termites. Waterborne wood preservatives are commonly used when the wood may be in contact with humans or will be painted.

Appendix B: Glossary of Terms A p p e n d i x B-G l o s s a r y o f T e r m s Fire Depletion Area Burned: Fire Impacts: Fire Intensity: Fire Load: Fire Management: Fire Management Zone: Fire Prevention: Fire Protection: Fire Regime: Fire Risk: Area burned that directly impacts wood supply to the forest industry. This could include allocated .

social or cultural context (livelihoods, festivals, traditional, conflict) and perhaps regulatory framework (permit fires, illegal fires). The terms include fires, wildfires, wildland fire, forest fire, grass fire, scrub fire, brush fire, bush fire, veldt fire, rural fire, vegetation fire and so on (IUFRO 2018). The European Forest Fire

Structural (1996), the 16-hour Fire Control 3A: Structural Fire Fighting in Acquired Structures (2009), and/or the 16-hour Fire Control 3B: Structural Fire Fighting in Live-fire Simulators curriculum. 4. The NFPA standards identify the requirement that all fire fighters shall demonstrate a

wood trusses. The versatility of wood trusses makes it an excellent roof framing system in hybrid construction where wood trusses are commonly used with steel, concrete or masonry wall systems. Environmental: Wood, the only renewable building material, has numerous environ-mental advantages. Wood trusses enhance wood's environmental

The Being a Writer program combines a writing process approach with guided instruction to ensure students learn and practice the craft and conventions of writing. Every lesson operates in the context of a caring classroom community, crucial to motivating and inspiring students to grow as writers, thinkers, and principled people. The program is built on the assumption that academic and social .