2 FUNDAMENTALS OF MACHINE DESIGN MIR PUBLISHERS MOSCOW P QRLOV n. opjiob OCHOBbI KOHCTPyHPOBAHHH H3,Il,ATEJIbCTBO «MAUIHHOCTPOEHHE» MOCKBA FUNDAMENTALS OF MACHINE DESIGN P. ORLOV TRANSLATED FROM THE RUSSIAN by YU. TRAVNICHEV MIR PUBLISHERS • MOSCOW First Published 1976 The Russian Alphabet and Transliteration A a a Kk K Xx kh B 6 b JI ji 1 An ts B B V M M m Hu ch r r g Hh n in m sh Ah d Oo 0 id, m shch E e e nn P T> t i 1 E e e Pp r LI H y JK ;k zh Cc s Lb * 3 3 z T T t 33 e M h i vy u K> io yu Hi y (D$ f Ha ya The Greek Alphabet A a Alpha 11 Iota Pp Rho Bp Beta Rx Kappa 2a Sigma r v Gamma Ak Lambda Tt Tau A 6 Delta M p. Mu r u Upsilon Ee Epsilon Nv Nu Phi z? Zeta BE Xi XX Chi H T) Eta Oo Omicron Psi ©0 Theta n n Pi Q (o Omega Ha amJiuucKOM H3UKe + n + h ) kgf-m (1.1) where P ax = axial force arising as a result of the bolt tightening, kgf d 0 — pitch diameter of the thread, mm D = mean diameter of nut bearing surface, mm 8 Chapter 1. Tightened Connections h and / 2 = coefficient of friction in the threads and on the nut bearing surface, respectively

where d is the nominal diameter of the thread, mm. For the current diameter ranges of fastening bolts it is usual to take on the average s/d = 0.15; d 0 /d = 0.9; and D/d = 1.3. Substituting these values into Eq. (1.2), we get Mtight = \0~ z P ax d (0.024 + 0.45/j + , + 0.65/ 2 ) kgf-m whence p _10 3 _ ^Jight - d (0.024+0.45/i + 0.65/ 2 ) (1.3) Let f 1 = 0.22 and / 2 =0.11. Then 5 M tight <1 do d, ■D $ 10 3 - d (1.4) Fig. 1. Determination of bolt tightening torque Force P ax induces tensile stresses in the bolt rr — ^ ax Ufens ' 0.785df where d x is the minor diameter of the thread (for light bolts the diameter of the bolt stem), mm. The moment of friction in the threads — fi causes torsional stresses in the bolt _ Pax^ofi 2W tors where W tOTB — 0.2 d\ is the torsion resisting moment of the bolt cross-section. Consequently _ P T ~ 0.4di* The total stress, according to the third theory of strength, is a^Vohns-r 1.6 + 25/? (1.5) 1.1. Unloaded Connections 9 Assuming (on the average) that = 0.8d, and substituting the values / x = 0.22 and d 0 /d = 0.9, we obtain 2.6 ^ d 2 ( 1 . 6 ) Substituting into this expression the value of P ax from Eq. (1.4), we have 10 3 - ^ gM kgf/mm 2 (1.7) where M tight = tightening torque, kgf*m d = nominal diameter of the thread, mm The values of a versus M tight , shown in Fig. 2, are estimated from- Eq. (1.7) for bolts of different diameters. The diagram can be used Fig. 2. Tightening torque Mti g ht an d stresses a for bolts of different diameters^ for approximate determination of stresses developing in a bolt tightened with various torques. From the permissible stress it is= possible to find the ultimate tightening torque. 10 Chapter 1. Tightened Connections The inverse-cube relationship between the stresses due to tighten¬ ing and bolt diameter [see Eq. (1.7)] accounts for the abrupt rise ■of the stresses, occurring when the bolt diameter is reduced. When manually tightening small-diameter bolts, one may overstress the holts, thus stretching and even breaking them. Approximate values of forces and torques when tightening bolts by hand are given in Table 1. Table 1 Forces and Torques when Tightening Bolts Manually Bolts Wrench arm, m Tightening force, kgf Tightening torque, kgf-m .Small (M4-M8) 0.1-0.15 » 10 1-1.5 Medium (M10-M14) 0.15-0.2 «15 2-3 Large (M16-M24) 0.2-0.25 «20 4-5 Assume, for example, that the tightening torque is 1.5 kgf •m. Drawing in Fig. 2 a horizontal line Mu g ht — 1.5 kgf-m, we read the stress values on the abscissa axis: for bolts M8—37 kgf/mm 2 ; for bolts M6—80 kgf/mm 2 . The latter figure is well above the yield limit of commercial carbon steels. Consequently, bolts smaller than ZM6 may easily be broken when tightened by hand, and with the application of excessive forces bolts M8 may also be broken. The magnitude of stresses due to tightening, in accordance with ZEq. (1.3), strongly depends on the coefficient of friction in the threads and on the nut bearing surface. Friction blocks, as it were, the tighten¬ ing force, so that a major part of the latter is spent in overcoming friction, only an insignificant portion being taken up by the bolt stem. For example, when f x = 0.22 and / 2 = 0.11 the portion of the torque utilized for the bolt tightening, according to Eq. (1.3), will be 0.024 0.024+0.1 + 0.72 • 100 % 12 % The remaining 88% of the torque is spent in overcoming friction. Equation (1.7) and the diagram in Fig. 2 are based on rather high values of the friction coefficients (f x = 0.22; / 2 = 0.11) correspond¬ ing to non-lubricated surfaces. If some lubricant gets onto the friction surfaces, then, with the same torque, the bolt stresses will rise. Table 2 gives the values of stresses, calculated from Eq. (1.7), -which arise in bolts tightened by means of standard wrenches with the application of 15-kgf tightening force. From the Table it is clear that with low friction values it is possible to break even M10 bolts -when tightening them manually. The possibility of overstres- 1.1. Unloaded Connections 11 * Table 2 Stresses in Bolts when Tightening with Standard Wrenches Bolt dia. j Stresses in kgf/mm 2 when fi = 0.22 / 2 = 0.11 /i = 0.11 h = 0.055 M6 100 180 M8 50 90 M10 30 54 M12 17 30 M14 12 22 M16 9 16 Note. The heavy line delimits stresses exceeding the yield limit of commercial carbon steels. sing bolts with threads larger than M12 when tightening them with standard wrenches is practically excluded. If, however, small-sized bolts must be used for some design reasons, then appropriate measures must be taken to limit the tightening Fig. 3. Methods of eliminating bolt twist when tightening torque or the bolts must be made from a high-grade heat-treated steel. The simplest way to limit the tightening torque is to shorten the wrench arm correspondingly with the decrease of the bolt diameter. This practice is, in effect, stipulated by current standards covering wrenches and spanners. 12 Chapter 1. Tightened Connections The twisting-off of a bolt can be avoided if during tightening it is held by a special element (Fig. 3 a) or if the bolt end is locked to the housing (Fig. 3 b). In these cases only tensile stresses occur in the bolt. Assume that in Eq. (1.5) r = 0, and dy = 0.8d (as before), then a = 1.56 ijs. a i Comparing this expression with Eq. (1.6), we see that the stresses 1 56 amount only to 0.65 of the stresses when tightening a bolt with its being twisted. Torsional stresses arise only during tightening and subsequently vanish due to the bolt elasticity. Therefore, when designing tightened connections for long-term strength the torsional stresses are usually neglected, the bolt calculations being limited to the axial force Pax only [see Eq. (1.3)]. 1.2. Loaded Connections These include connections subjected to the action of a force which stretches the joint and imposes an additional load upon the fastening bolts. This force can be constant (e.g., the static pressure of gases and liquids in vessels) or variable (the pressure of gases in internal combustion engines and piston compressors, inertia forces of moving masses in connecting rods and bearings of crank mechanisms). Here the bolt preload is selected so that, regardless of possible working force fluctua¬ tions, a constant tension at the joint is maintained, avert¬ ing its opening which could break the joint seal and, in the case of variable forces, cause the crushing and work hard¬ ening of the metal surfaces. The connection may be additionally loaded by thermal forces aris¬ ing as the system gets heated. Figure 4 shows a bolted connection subjected to the action of internal working pressure P work . To assure normal functioning of the joint the bolts must be pretightened by a force P tight sufficient to keep the joint tight under the action of working force P wor k- Now let us clarify what deformations occur in the system when force Ptight is applied to it. For the sake of simplicity we neglect changes 12. Loaded Connections 13 in the length of the threaded bolt ends and consider that the working length l of the bolts is equal to the thickness of the joined parts. Under the action of force Ptight the bolts will lengthen by the value , Plight ** ~ EiFi while the joined flanges will shrink by the value , Plight E 2 F 2 where E i, E 2 and F 2 , F 2 are the elasticity moduli of the materials and cross-sectional areas of the bolts and housing, respectively. The force with which the housing members are compressed is equal to the tightening force, i.e., Pcomp = P tight (1-8) After the application of the working force P wor h the bolts will additionally extend by the value AX, and, accordingly, the housing compressive deformation will decrease by the same value. As a result, the pressure exerted by the housing upon the bolts is reduced by the amount A P. The force tensioning the bolts becomes P tens — Pwork "l” Ptight &P (1 -9) and the force of the housing compression P comp = P tight — &P (1-10) Force A P can be deduced from the following relationships. In accord with Hooke’s law, the value of the housing deformation decrease is In bolts the same amount of deformation is produced by the dif¬ ference of forces prior to and after the application of force P w0Th , i.e., Ptens Ptight = Pwork -f* Ptight AP Ptight = Pwork A P Consequently, in respect to the bolts AX (Pworh AP) l EiF j Equating Eqs. (1.11) and (1.12), gives A P Pwork 1 E,F i E 2 F 2 ( 1 . 12 ) (1.13) 14 Chapter 1. Tightened Connections Substituting this expression into Eqs. (1.9) and (1.10) gives the bolt tension force Ptens = Pwor k + Ptitht - A Pw °[ h fV = Ptitht + Pw £$ - 2 ( 1 - 14 ) 1+ E 2 F 2 l+ EiFi and the joint compression force Pcomp ~ Ptight -( 1 - 15 ) If the working force varies from zero up to P worh , then the bolt tension force will pulsate with an amplitude At ens = Ptens- Ptight = Pw ° T f- (1.16) 1-1-LA + E\F \ while the housing compression force will pulsate with an amplitude *comp • = Ptight—P‘ work comp ' EjFj E 2 F 2 (1.17) Length does not enter into Eqs. (1.14)-(1.17). This means that the forces acting at the joint are theoretically equal when tightening low flanges and high housing components by bolts. Actually the forces under consideration are affected by the elastic and plastic deformations of the threads and bearing surfaces of the nuts and bolt heads, etc., which may appreciably reduce the forces that tension the bolts and compress the connection. The relative significance of terminal conditions is much greater for shorter bolts than for longer ones. That is why the joints fastened with short bolts weaken more quickly in use, particularly when sub¬ jected to pulsating loads. As a general rule, it is better to use long bolts (high flanges) or introduce elastic elements compensating for local plastic deformations of the system. On the basis of Eqs. (1.14) and (1.15) it is often concluded that smaller ratios E 1 F 1 /E 2 F 2 are more beneficial, in other words, the best combination is elastic bolts and rigid housings. From Eq. (1.14) it is evident that the bolt tension force is minimum E F (Pt ens = Ptight) when E ' ~p - = 0 (perfectly elastic bolts or perfectly ^ ^ Fa F rigid housings) and increases with an increase in ~r -, reaching 2* 2 E F its maximum (P ten s = Ptight + Pworh) when ^r- = °o (per¬ fectly elastic housings or perfectly rigid bolts). The amplitude of the pulsating tension force [Eq. (1.16)] also falls down with a decrease in tt-tt- • In the limiting case (~~ = 0) the bolt tension force 2 / 1.2. Loaded Connections 15 remains constant and is Ptens = Ptight > i.e., the load on the bolts becomes static despite the pulsating working force.-With .an increase in ~^r- the load on the bolts becomes cyclic. When — oo A 2^2 £>2*2 the pulsation amplitude is A fens = P W0T h- J? F With a decrease in g~ l the joint compression force P CO mp [Eq. (1.15)] is also reduced. This is beneficial for the housing strength, but is disadvantageous for the joint sealing, since the force sealing the split connection is equal to P comp . At the same time, with the decrease of - ~iS the pulsation of force P CO mp [Eq. (1.17)] is increas- ■^ 2^2 ed, but since the housing strength is usually much greater than the bolt strength and the compression force fluctuations are not so dangerous as the bolt tension force fluctuations, then it is recommend- ed to employ low ^ values, as it is considered that this lowers £-2^2 the bolt load. From this stems the time-honoured design rule for split connections: elastic bolts—rigid flanges. To this rule certain essential corrections are necessary. To fully reveal the essence of this phenomenon, it is necessary to ascribe definiteness to the member Ptight entering into Eqs. (1.14) and (1.15), i.e., to agree upon how to select the tightening force. There are two methods of such a selection. In accord with the first method, widely applied until recently, the tightening force was taken in proportion to the working force P W0T k Ptight— yP work (1.18) where y is the tightening factor (usually y = 1-2). According to the second method, the tightening force is found on condition that the joint compression force P co mp is proportional to the working force, i.e., P comp — QP work (1.19) where 9 is the proportionality factor (generally 0 = 0.25-1). Let us consider both of these cases. The case when Ptight — yPwork■ Substituting into Eqs (1.14 and 1.15) Ptight = yPwork, gives Ptens — Pwork (y-\ -( 1 . 20 ) V 1+ ^7 / Pcomp —Pworh ^7 EtF\ ^ (1.21) \ EzPz ) F P From Eq. (1.20) it is clear that variations of the ratio even within the widest limits affects the magnitude of the force Ptens 16 Chapter 1. Tightened Connections comparatively little. In the limiting cases Ptens = Pwork! (when |= 0) and P tens = P WOTh (1 + y) (when = oo) . Hence, the entire range of force Ptens is confined within the limits . TT P It should be noted that in the expression ■ - 1 the number of •& 2^2 possible values of the ratio EJE 2 is limited. Bolts are made almost fig. 5. Ratios P tens /P work and Pcomp 1 p work as a function of F 1 /F 2 for diffe¬ rent values of tightening factor y exclusively of steels (E = 20"lO 3 - 22-lO 3 kgf/mm 2 ) and in very rare cases (in special constructions) — of titanium alloys (E = = 11.5-10 3 -12.5-lO 3 kgf/mm 2 ). Joined members are made of steels, cast irons, light and titanium alloys (Table 3). There are three values of the ratio E r IE 2 having practical impor¬ tance: EJEz — 1 (steel-steel; titanium alloy-titanium al\oy)',Ei/E 2 w « 2.5 (steel-cast iron); EJE^ w 3 (steel-aluminium alloys). With specified materials the value can be influenced only i by changing the ratio FJF 2 , which entails changes in the strength of the bolts and housings. When clamping a split aluminium housing with steel bolts (EJE 2 w 3) the ratio Ptens/Pwork, for different values of y taken in a wide range of FJF^ values (from zero to unity), will vary inap¬ preciably, only by 1.5-2 times on the average, as evident from Fig. 5a. 1.2. Loaded Connections 17 , Table 3 Material Combinations for Bolts and Joined Components Bolt material Joined components Ei/E% material ® 2 , kgt/min! Steel Steel i Ei =21000 kgf/mm 2 Cast iron 2.6 Aluminium alloys 2.9 Magnesium alloys 4 500 4.7 Titanium alloys 12 000 1.75 Titanium alloy Titanium alloys 12 000 1 Ei = 12 000 kgf/mm 2 Thus, the gain from decreasing F 1 /F 2 is not very great. The picture is similar for other EJE 2 ratios. Decreasing the force Pt en s still does not mean an improvement in the bolt strength which is determined by the stress a = . Such an improvement will only be reached if the ratio FJF 2 is reduced by increasing the housing cross-section and not by reducing the bolt cross-sections. It can easily be proved that reducing the bolt cross-sections decreases relatively little the force Pt enS i and, at the same time, sharply increases the bolt stresses. On the other hand, increasing the ratio FJF 2 by enlarging the bolt cross-sections provides a definite gain in the bolt strength. In relation to joints for which good sealing is imperative it is extremely important to find how the compression force P CO mp changes with F x !F 2 , for this force determines the tightening of the joint seal. The values of the ratio Pcomp/Pwork (for EJE 2 = 3), estimated from Eq. (1.21), are plotted in Fig. 5b. The compression force decreas¬ es with the reduction of the ratio F x /F 2 (rigid housings), the decrease being the sharper the less the value of y. For example, with y = 1.25 force P C omp within the range of FJF 2 of from 0 to 1 is reduced four times. Hence, the decrease of the ratio FJF 2 adversely affects the reliability of the joint. To assure reliable sealing in the case of rigid housings it is neces¬ sary to increase the tightening force, i.e., to increase the bolt tension, which makes the gain from lowering F X !F 2 fictive. Conversely, making the housing more pliable lowers the necessary tightening force. With compliant housings it is possible to tighten 2-01418 18 Chapter 1. Tightened Connections the joint with y < 1 without running the risk of the joint loosening ( Pcomp/P work > 0) and without impairing its reliability. Let us now examine the effect of the ratio FJF^ upon the pulsation amplitude of forces Pt ens and P CO m P (Fig. 6). It is evident from the graph that decreasing the ratio FJF^ (rigid housings) lowers the pulsation amplitude of force P tens , which is advatageous for the bolt strength. However, the pulsation of force P com p increases as the ratio Fi/.F 2 is reduced, this adversely affecting the reliability of the joint. Thus, we may conclude that with P tirht = yP WO rh, low ratios FJF t (rigid housings) are advantageous for the bolt strength when Fig. 6. Ratios A tenJPwork and &comp/ p worh as a function of Fi/F 2 for diffe¬ rent values of tightening factor y the loads on the joint are pulsating. For static (in particular thermal) loads, and also when the reliability of the joint is of prime importan¬ ce, higher ratios FJF 2 (compliant housings) are preferable. The foregoing conclusions are of special significance for connect¬ ions in which the rigidity of the joined parts is commensurable with that of the clamping bolts. Such a situation is encountered, for instance, when clamping hydraulic and air power cylinders (Fig. la) and cylinders of internal combustion engines and piston compressors of half-block (Fig. lb) or full-block (Fig. 7c) designs. Here the design¬ er can vary within broad limits the ratio F-JF 2 by changing the cross-sections of the clamped components and thus make it most suitable for the given operating conditions. With conventional flange connections manoeuvring possibilities are less. As shown by experiments, actually in operation is the cylin- 1.2. Loaded Gonnicfions- 19: drical flange volume of external diameter D « 2 d 0 and internal diameter d « d 0 (Fig. 8a). The ratio of the bolt cross-sectional Fig. 7. Tightened connections area to the design area of the clamped parts is F z ~ D 2 -d% This ratio can be increased by placing an elastic gasket between the parts being connected, or reduced by placing under the bolt some massive washers of '~d H increased diameterjand also l"~ j°~j i i °~j by decreasing the bolt stem ( r I 7 1 (j tS diameter (Fig. 8b). I I I ( J -*— Assuming that D x = 3d„ and d x = 0.8 d 0 , we have ///fcjf Ikxaa FaIxx 1 1xXavv' (0.8d 0 ) 2 = 0.08 The ratio F x /F 2 can also \\\K< | X\v vf aona; be reduced by placing ela- aw \v\jgg_._j) . ggcw stic elements between the Aw Sw t \ \ toa flange and bolt faces. . vv p al.,. . — Avt ori l l | .H &gj w The case when P comp = I 111 'l I ! I \ =QP wor k- The relationships ' / ) ^fb) become different when the ' a ' tightening force is chosen Fig. 8. Tightened flanges so that Pcomp of the joint is proportional to P worh . This condition is quite logical: the higher the working pressure, the greater must be! the compression to ensure the required reliability of the'joint. Selecting the' tightening force in such a way establishes a direct relation between the joint com-? to Chapter 1. Tightened Connections pression force and the working force, whereas in the previous case this halation is a derivative of the preliminarily chosen tightening factor y and of elasticity characteristics of the system. Thejbolt tension force is always equal to the sum of the working and compression forces. Hence, Ptens — Pwork Pcomp — (1 “l - 6) Pwork (1.22) Tjhus, the forces P tens and P comp (for a given force P worh ) in this case are constant and independent of the ratio E^FJE^F^. From Eq. (1.15) the required tightening force Pwork , ElFi (1.23) Substituting into this relation the value P comp = QP worh , we get Ptight ~ Pwork (Q “I-\ (1-24) The ratio PtightIPworh as a function of FJF 2 (when E X IE 2 = 3) is plotted for various values of 0 in Fig. 9. The graph shows that p increasing the ratio FJF 2 ^tiqht (compliant housings) redu¬ ces the required tightening force. The determination of the tightening force on condi¬ tion that Pcomp — 0 Pwork is undoubtedly more ration¬ al than on condition that = yP, The lat- . 9. Ratio Ptight!Pwork as a function of Fi/F 2 for different values of factor 0 ter method must be reject¬ ed as being wrong in prin¬ ciple. In this connection the problem of the influen¬ ce of the E 1 F 1 /E 2 F 2 factor on the operation of the jo¬ int should be reconsidered. As is obvious from the fore¬ going, when Pcomp—QPwork the ratio E 1 F 1 /E 2 F 2 has no influence on the forces Ptens and Pcomp which are de¬ fined exclusively by the magnitude of factor 0. The ratio EiFi/E 2 F 2 affects the tightening force only. Higher values of EiF 1 /E 2 F 2 (compliant housings) are more advantageous as they enable the required tightening force to be decreased. 1.2. Loaded Connections 21 In connections loaded by a pulsating force the ratio EiFi!E 2 F 2 also affects the pulsation amplitude of forces Ptens and P com p. According to Eqs. (1.16) and (1.17), as E X F X !E 2 F 2 decreases (rigid housings), the pulsation amplitude of Pt ena is reduced, while that of Pcomp, increased. 1 Conversely, increasing E 1 F 1 /E 2 F 2 (compliant housings) augments the pulsation amplitude of Pt en s and diminishes that of Pcomp• Consequently, we may conclude that under pulsating loads rigid housings are more advantageous for the bolt strength and compliant housings, for reliable sealing. In joints loaded by a constant force the magnitude of E 1 F 1 /E i F„ seems indifferent both for the bolt strength and for the quality of sealing. All the above considerations on the relative advantages of rigid and compliant housings under a static load are valid only when P tight = yPwork and lose their value when P eo mp = QPworh- The magnitude of factor 9, which in the given case is of decisive importance for the joint parameters, is chosen to suit the reliability requirements for the seal: for non-critical joints .6 = 0.25 to 0.3 and for critical joints, 0.5 to 1. The calculation on condition that P CO mp = QPworh is simpler. There is no need for the cut-and-try procedure of finding Ptight and checking the resulting P co mp, and the calculation comes to the application of the simplest formulae (1.22), (1.23) and (1.24) which at once yield all the values determining the strength and reliability of the joint. (a) Temperature Factors If a joint operates at elevated temperatures, and the temperatures of the clamping bolts and joined parts differ, or if the parts are made of materials possessing different coefficients of linear expansion, a thermal stress P t then develops in the joint. In accord with Eq. (7.1) [see Fundamentals of Machine Design , Vol. 1, Chapter 7], this stress is Pt = (a 2 * 2 — / n p p Pwork i (^2*2 ~ °Ml) ,a , - d- 28 ) Introducing this factor into the previous equations in place of E 2 F 2 , the calculation can be carried out as previously. In this case the temperature deformation is Ae = + Va’Jl + ... — la^ where a 2 = coefficient of linear expansion of the bolt material ti = bolt temperature In relative units Ae=*L a '/ 2 +I-a;t;+...- ai t l (1.29) The sum of the deformations of the individual elements in the system may be expressed as a function of the thermal stress in the following way: Ae-- tPl' P t l" E'F'l ' E’F’l 1 • ' • 1 E^t Equating Eqs. (1.29) and (1.30), one will obtain ~i —b a a*5 “+'••• —F i — p * ( eg + EiFi) Pt= EjFj eg (1.31) (c) Elastic Elements The elasticity of a bolt-housing system can be varied without changing the sections of the system components by introducting elastic elements (Fig. 106). This is widely employed in practice. Depending on their arrangement, such elastic elements increase the elasticity of either the bolts or housings. In order to evaluate the effect of the elastic elements it is necessary first to find out which elements belong to the bolt system and which to the housing one. If the application of the working force P WO rh increases the load on an element, the latter then belongs to the bolt system, regardless of whether the load is tensile or compressive. If the working force lessens the load on the element, this element then belongs to the housing system. 24 Chapter 1. Tightened Connections For instance, the load on element 1 (Fig. 10b) is increased when the working force P WO rk Is applied. Therefore, this element belongs to the bolt system; its elasticity should be included into that of the fastening bolts. Element 2 , on the contrary, is unloaded when the working force is applied and, hence, must be regarded as belonging to the housing system. The deformation of an elastic element under the effect of the applied load is V = Pfl. where l e = length of the element / = elastic ratio of the element (relative deformation pro¬ duced by 1-kgf force) The deformation related to the total length l of the joint is e' = Pf l f The total relative deformation of a bolt with an elastic element having length l a and elasticity f a is the sum of the deformations of the bolt and element e = e 1 + e ' = p(-JL The quantity 1 EiFi P (1.32) is the corrected rigidity factor of the system. The corrected rigidity factor of a housing with an elastic element of length l b and elasticity f b is ® 6 & 6 =-T - J —IT (1 - 33) ~E^ +fb T These quantities can be introduced into the previous equations in place of E 1 F 1 and E 2 F 2 and the calculations carried out as pre¬ viously. With the help of elastic elements the operating parameters of a connection can be beneficially influenced. For example, intro¬ ducing elastic elements into the housing system can improve the reliability of sealing of the connection and reduce the necessary bolt tightening force. The pulsation of the bolt tension force can also be reduced by introducing elastic elements. Elastic elements are an efficient means for preventing the gradual weakening of the tightening force due to relaxation. 1.2. Loaded Connections 25 ( d) Relaxation Tightened joints (particularly those operating under elevated temperatures) in the course of time weaken due to a slowly deve¬ loping plastic deformation of the bolts (and, occasionally, of the clamped parts as well) under the influence of loads acting for a long time. The plastic deformation phenomenon occurring under stresses well below those corresponding to the yield point of the material is called relaxation. The property of metals to flow when being acted upon by a continuous load can be ascertained only in special tests, during which specimens are held under load for 3000 to 10 000 hours. Relaxation is often defined as a spontaneous change with time of stresses with unchanging deformation. This definition cannot be accepted. The fall in stresses during relaxation is inevitably accom¬ panied by plastic deformation. Moreover, plastic deformation is the first reason for relaxation. It is more legitimate to speak about a room-temperature creep of materials, which is akin to high-tem- perature creep, the difference being that the room-temperature creep deformation develops slower and has a smaller magnitude. The process of relaxation can be followed schematically on an example of housing-type components clamped with bolts. For the sake of simplicity we assume at first that the housing under consi¬ deration is perfectly rigid and the only deformable element is the holt. At first glance it seems that the system operates under the conditions of constant deformation. Actually this is not true. Even¬ tually the bolt elongates plastically. The original relative elastic deformation Ae of the bolt, due to the bolt preload, now decreases; by the amount of relative residual deformation A res so that the resultant elastic deformation becomes Ae' = Ae — A res . If the initial tightening force equalled Ptight — A eE x F x (E x — — elasticity modulus of the bolt material, F 1 = bolt cross-section), then after elongation the force becomes P'ught = (Ae — A res ) E X F X , i.e., it decreases in comparison to the original tightening force by (1 — times. Upon removing the bolt and measuring its length, it is found that the bolt has lengthened by the value lA rea (l — = original bolt length). With the decrease of the tightening force the stresses in the bolt fall down. When they reach the level at which the bolt elongation discontinues or becomes negligible, the process of relaxation ceases and the stresses in the system become stable. In real systems of clamped elastic parts the course of this process is somewhat different. As the bolt elongation proceeds, the clamped parts, while expanding elastically continue to exert pressure on the bolt, although somewhat lower in comparison with the initial 26 Chapter 1. Tightened Connections pressure. Under such circumstances the relaxation process stops and the system becomes stable with a relatively greater bolt elongation than in the previous case. We have analysed a non-loaded connection. In systems subjected to a constantly acting static or cyclic load relaxation progresses continuously. The state of practical stabilization sets in only as a result of the slowing down of the elongation process, which is •observed in most materials in the nection has not failed before this Fig. ila. Total deformation e of a car¬ bon-steel specimen (0.1%C) after being held under tensile stress o. Test temperature 20°C (after Rogan Alexander) course of time, provided the con- due to the separation of the joint Fig. 116. Residual stresses in per cent of original stress (o 0 = 25 kgf/mm 2 ) as a function of holding time. Test temperature 400-450°C faces. The higher the stress, the greater the creep of materials ■(Fig. 11a), and it sharply increases with the rise in temperature. The creep is greater under cyclic loads than under static ones. It varies with different metals and alloys and depends on the type of heat treatment. This requires the introduction of a concept of relaxation resistance understood as the ability of a material to resist plastic deformation mnder continuous loads. The method of determining relaxation resistance is not as yet fully -established. The majority of the methods applied at present are based on inducing in specimens a certain, rather high stress; the specimens are held at this particular stress and at an elevated tern- 1.2. Loaded Connections 27 perature for a long period of time (up to 10 000 hours). Generally, the chosen stress amounts to 0.5-0.6 of the yield limit of the material at the given temperature. A split-ring method is the one most often employed. The specimen is a ring with a wedge-shaped cut. The required stress is produced by forcing a wedge into the cut, which causes bending stresses to develop in the opposite ring section shaped so as to have the same bending resistance throughout its entire length. In this state the specimen is kept at the given temperature for the given period of time. At definite time intervals the load is removed and the residual deformation measured. This is used for calculating the stress remain¬ ing in the specimen at the end of each interval. The diagram shown in Fig. 11 is a result of such tests. The re¬ maining stresses are shown in Fig. 11 b as a function of time and characterize the relaxation resistance of the material. The higher the stress remaining in the specimen (i.e., the lower the residual deformation of the specimen), the greater the relaxation resistance of the material. The remaining stresses sharply drop during the first 1000 hours of tests and then lower at a slower pace. This shows that as the holding period grows longer the yielding of the material decreases (seemingly as a result of strain hardening). The relaxation resistances of various materials are different. Thus, for example, a specimen made from steel grade 40X, after being held at 400°C for 3000 hours, retains 80% of its initial stresses, while one made from steel grade 50XA retains only 40%. If we convert the results obtained for bending loads to the case of tensile loads, then a bolt made from steel 50X0A will, after 3000 hours’ work, extend to a length equal to 60% of its initial elastic deforma¬ tion due to tightening, in which case the tightening force will be reduced by 60%. Naturally, account should also be taken of the above-described effects of the elasticity of the clamped parts and of the working forces which add to the bolt elongation. Estimating the relaxation resistance by the magnitude of the residual stresses is controversial. To reflect the physical essence of the phenomenon and make calculations more convenient it is advis¬ able to characterize the resistance to relaxation by the magnitude of the deformations remaining in the specimen after its being held under a load corresponding to the actual loading conditions (for fastening bolts, under a tensile load). Methods to prevent the weakening of tightened connections as a result of relaxation consist in making bolts from relaxation-re¬ sistant materials and subjecting them to appropriate heat treatment. Silicon steels possess a high relaxation resistance. Normalizing followed by high tempering is an optimum heat treatment from the standpoint of relaxation resistance. It seems possible to obtain 28 Chapter 1. Tightened Connections a significant effect by strengthening the bolts through their preli¬ minary elongation (ageing) which acceleratively simulates the greatest yield stage of the material. All measures should be taken to reduce stresses in bolts and the pulsation amplitude of the tension force in cyclically loaded con¬ nections. It is also advisable to introduce elastic elements of relaxa¬ tion-resistant materials so as to compensate for residual deforma¬ tions as they develop. (e) Graphical Calculations of Tightened Connections The totality of the phenomena occurring in tightened connections is easily amenable to graphical interpretation with the help of P-e (force vs. relative deformation) diagrams. Let us take the simplest case, namely, the tightening of a split connection with force Pu g ht (Fig- 12a). In solving the problem the forces will be plotted on the ordinate axis, and relative deforma¬ tions — on the abscissa axis, assuming tension to be a positive, and compression, a negative deformation. Line Oa shows the bolt elongation and its slope relative to the abscissa axis is tan a = — EiF ± where T] = scale of forces fi = scale of relative deformations The compression of the parts being tightened is presented by the straight line Ob whose slope relative to the abscissa axis is tan P — -jj- E 2 F 2 If we draw on the diagram a horizontal line ba at a distance cor¬ responding to the tightening force Pttght from the abscissa axis, it will intersect the tension-compression lines at points a and b, the abscissae of which are equal to the relative deformations e 1 of the bolts and e 2 of the housing when tightened. It is more convenient to plot the diagram as shown in Fig. 12 b, i.e., by drawing the compression line through point a on the tension line, which corresponds to P = P tight- Now assume that a force P W0T k develops in the joint and that this force imposes an additional load on the bolts while unloading the joint (Fig. 12c). When determining the bolt tension force, we assume, as previously, that the relative deformations of the bolts and clamped components are the same. At a distance P W0T h (along the vertical) from the bolt tension line Oa draw a straight line (dashed) parallel to it. Through point k. jvtyd - 1.2. Loaded Connections 29 where this line intersects the clamped parts’ compression line, draw a vertical line until the latter intersects the tension line (point c). This vertical line will cut off a line-segment of length Ae on the Fig. 12. Graphical calculations of tightened connections abscissa axis. Clearly, this scheme reproduces the condition of equal Ae for the bolts and clamped components. This construction gives all the necessary data. The ordinate of point c shows to scale the bolt tension force Pt ens ', the ordinate of point k, the joint compression force P com p\ line-segments e[ and e , 30 Chapter 1. Tightened Connections relative deformations of the bolts and clamped components after application of working force P wor k- If the working force pulsates from zero to P w0Tk , the pulsation amplitude of force P tens will then equal Ptens — Pttght (the right- hand wavy curve) and that of force P comp , Pttght — Pcomp (the left-hand wavy curve). Now let us consider how the system’s behaviour is influenced by the variation of the rigidity of bolts and clamped parts (i.e., by changes in the slope angles a and ft of the tension and compression lines, respectively). Figure 12 d illustrates the case of a rigid housing and elastic bolts, while Fig. 12e, that of rigid bolts and an elastic housing. The tighten¬ ing force Pttght is assumed to be the same in both cases. Figures 12 d and 12e clearly illustrate the above-described regularities. The increase of the bolt elasticity with a simultaneous increase in the housing rigidity (reduction of the ratio E^FJE^F^) lowers the bolt tension force and the pulsation amplitude of this force (see Fig. 12 d). At the same time this decreases the joint compression force and in¬ creases its pulsation amplitude. Increasing the ratio EtF^E^F^ (elastic housing and rigid bolts) acts in the opposite direction (see Fig. 12e): force Ptens and the amplitude of its pulsation increase, P CO mp rises too, but the ampi- tude of its pulsation decreases. As stated earlier, when determining connection parameters it is more proper to proceed from the condition of proportionality between the compression force and working pressure (P comp = = 6P work)- In this case the graph is plotted in the following order (Fig. 12/). First, draw horizontals at distances P CO mv = QPwork and Ptens = = (1 + 0) P WO rk from the abscissa axis (in Fig. 12/ it is assumed that 0 = 0.6). Then from any arbitrary point draw at an angle (1 an inclined line representing the housing compression. From point d at which this line intersects the horizontal line P comp draw a ver¬ tical until it meets the horizontal line Ptens and through point e of their intersection draw at an angle a an inclined line representing the bolt tension. The ordinate of point /, where the two inclined lines intersect, gives the value of the tightening force Pttght- The difference Ptens — Pttght shows, as before, pulsation ampli¬ tude of force Ptens , and the difference Ptight — Pcomp i the pulsation amplitude of force P comp . Figures 12 g and 12 h show how the rigidity of the housing and bolts influences the operation of the joint (with P worh and 0 values being the same as above). In the case of elastic bolts and a rigid housing (Fig. 12g) force Pttght increases, the pulsation amplitude of force Ptens decreases and the pulsation amplitude of force P com p increases. 1,2. Loaded Connections 31 In the case of rigid bolts and an elastic housing (Fig. 12 h) the force Ptight decreases, the pulsation amplitude of force Ptens increas¬ es and the pulsation amplitude of force P CO mp decreases. For an unloaded connection (P WO rk = 0) subjected to heating, which causes a thermal stress P t to develop in the joint (Fig. 12i) r one should draw a line parallel to and spaced from the compression line at a distance of a 2 t 2 — a 1 t 1 (along the horizontal), equal to the relative deformation due to the heating. The ordinate of point a', at which this line intersects the tension line, is equal to Ptens (iden¬ tical to force Pcomp compressing the housing) after the heating; the difference between the ordinates of points a and a’ represents the thermal stress P t . For connections which are simultaneously subjected to heating and to the action of force P WOT k the plotting differs in that the inclin¬ ed tension and compression lines are set apart (horizontally) at a distance of a 2 t 2 — a 1 t l (Fig. 12;). The rest of the construction is as before. The elasticity characteristic of the clamped parts does not always follow the linear relation P = eE 2 F 2 . Such is, for instance, the case with thin-walled intricately shaped components whose walls in certain areas are either inclined or perpendicular to the direction of the compressive force. Here the compressive deformations of the walls are complemented with the elastic bending deformations of their inclined and horizontal portions. Furthermore, the walls may be subject to buckling. If so, the rigidity of the clamped parts sharply drops and the characteristic curve assumes a gently sloping shape. Occasionally the elasticity characteristic becomes curvilinear because the individual elements of a construction come into opera¬ tion not at the same time, but one after another as the load increases. The elasticity characteristic of such complex constructions can only be determined experimentally. To this end a fully finished construction is subjected to compression on a test rig and its elasti¬ city characteristic is plotted point by point on the basis of the test data. In this instance the graphical method of calculation is the only way possible. The curve obtained from the tept is plotted on the graph in place of the inclined line P = eE 2 F 2 (Fig. 12 k, l). Further construction is as described above. When calculating for a known force P CO mp = QPworki the curvi¬ linear nature of the characteristic has an effect on the magnitude of the tightening force and of the pulsation amplitude of forces Ptens and Pcomp • With a convex characteristic (Fig. 12 k) the required force Pttght diminishes, the pulsation amplitude of force Ptens increases and that of force P CO mp decreases. If the characteristic is concave (Fig. 121) the picture is reversed. 32 Chapter 1. Tightened Connections (/) Methods of Controlling the Tightening Force {Preload) Because the preload has a great effect on the bolt tension and joint compression forces, it is very important during assembly to strictly Jceep to the calculated magnitude of the tightening force. This is •achieved by tightening nuts with torque wrenches, by turning them through calculated angles or by measuring the elongation of bolts. The application of torque measuring wrenches does not provide for adequate accuracy, since the force needed to tighten nuts depends to a large extent on the thread condition, coefficient of friction in the threads and on the bearing surfaces, etc. For this reason, bolts tightened with the same torque may, in fact, be loaded differently. To lessen the friction effect critical connections are sometimes tightened on a vibration table. The rather sharp decrease in the friction forces because of vibration must be considered when choosing the design tightening torque. When tightening nuts by rotating them through a calculated angle they are first brought into close contact with the bearing sur¬ faces, i.e., up to the point when bolts begin to extend. After this the nuts are turned with a wrench through the required angle v. This angle is determined from the given tightening force Ptight , proceeding from the following considerations. When tightening a nut, it is necessary to choose the bolt elonga¬ tion = P g 8 p ■ and clamped parts contraction X % = P ^ S p - . The axial displacement of the nut relative to the bolt is X = X z = Pught 1 (-e~fT+~e^) ( 1>34 ) The tangent of the thread helix angle tan( P = i < L35 ) where s — thread pitch d 0 = thread pitch diameter The displacement of the nut by value X corresponds to its rotation over arc C of the circumference of radius dj 2 C % tan q> (1.36) On the other hand, the value of C is C = iLv_4-.^V> (1.37) where v is the nut rotation angle in radians, and v°, the same in degrees. 1.2. Loaded Connections 33 Equating Eqs. (1.36) and (1.37), we have o X 360° t3Il (p 7t£?Q Substituting into this equation % and tan from Eqs. (1.34) and (1.35), we have v° = 360 Pught — (1.38) Clearly, the tightening angle is independent of the bolt diameter. Nuts are practically tightened by the following method. First all the clearances in the system are taken up (by tightening all the nuts home in a certain succession). This done, the nuts are slackened off and again tight¬ ened by hand or by means of a light torque-control wrench un¬ til the nuts are in close contact with the bearing surfaces. The¬ reafter all the nuts are tightened in a definite sequence (staggered, cross-like, serpentine-like) de¬ pending on the configuration of the joint faces and assuring a 4 uni- form tightening of the connec¬ tion by turning them first through an angle of v/2 and then, in the same succession, through the full angle v. To measure the nut rotation angle the wrench is provid¬ ed with pointer and the tight¬ ened component, a graduated disk. This method is more accurate than the first one, although it also contains sources of errors due to the difficulty of determining the right moment to start the tightening. It is only by fitting experience that this error can be reduced to minimum. In respect to slender bolts and studs the accuracy of measurement is influenced by their twisting during tightening, which is due to friction in the threads. To eliminate the twisting, the bolt end is held with a spanner when tightening (see Fig. 3a). However, this method is not convenient in actual fitting. Figure 13 depicts a wrench design which excludes any effect of the bolt twisting upon the accu¬ racy of the tightening angle measurement. Inserted into the wrench stem is a spring-loaded locking device 1 having a tapered end with a tapered shank of cross-like section. Fig. 13. Wrench design enabling ac¬ curate measurement of nut turning angle relative to bolt 3-01418 34 Chapter 1. Tightened Connections which fits into a corresponding socket in the bolt end. The locking device is spline-connected with rod 2 whose projecting end holds a friction-mounted pointer 3 positioned over graduated dial 4 attach¬ ed to the wrench end face. As the wrench is applied to the nut, the locking device enters the bolt socket, thus providing a direct connection between the bolt and pointer. Prior to tightening, the pointer is set to the dial zere position. During the tightening the pointer indicates the rotation angle of the nut relative to the bolt, i.e., the nut tightening angle. The most accurate method is that whereby the bolt elongation, i.e., % = * is measured directly. The elongation of short bolts is measured by means of a micrometer cal¬ liper gauge. The technique is appli¬ cable whenever the calliper jaws can be directly applied to the bolt ends. In this way, for example, is measured the elongation of connecting rod bolts, pinch bolts in clamp connections, fastening bolts of hydraulic cylin¬ ders, etc. For greater measuring convenience the bolt head and stem end are made spherical. The elongation of studs can be measured with the aid of an indica¬ tor, if its foot is set up on an independent support. When installed on the component being tightened the indicator will read the total of the elongation of the stud and the contraction of the tightened parts. Sometimes the tightening force is controlled by means of a system of deformable washers (Fig. 14). Rigid washers 1 and 2 and measuring ring 3 of a plastic metal (e.g., annealled copper) are placed under the nut. Reference washer 4 is placed concentrically around the measuring ring. The thickness of ring 3 is selected so that after the’ preliminary, light tightening of the nut a design clearance e (equal to the sum of elastic deformations in the tightened system) remains between the ring and reference washer. As the nut is finally tightened the measuring ring is flattened out. The tightening is stopped when the clearance e is taken up completely (this being indicated by the reference washer losing its mobility). The measuring ring must be changed whenever the nut has to be re tightened. (g) Calculation Example An engine cylinder block having a cross-section as shown in Fig. 15 is fasten¬ ed to the crankcase by means of studs 400 mm long; the stud stem diameter is 18 nun, thread M24 (pitch s = 1.5 mm). Fig. 14. Control of tightening force by means of deformable ring 1.2. Loaded Connections 35 Assume that the ignition force P u , orh = 10 000 kgf is taken up by the four studs nearest to the cylinder and that the studs' tightening force spreads over 0 0 Fig. 15. Cross-sectional view of cylinder block the block portion adjacent to the cylinder (delimited by lines 0-0 in the Figure) with a cross-sectional area of 7000 mm 2 . The block is made from aluminium alloy grade AJI5 ( E 2 = 7500 kgf/mm 2 , a 2 = 22 • 10 -6 ); the studs are of steel grade 30XTC (E 1 — 21 000 kgf/mm 2 ; = 11 - 10 _fl ). The temperature of the block and studs in the running engine is 80°C. Find the maximum stresses in the studs and crankcase in both the cold (dead) and hot (running) engine. Assume that 0 = 0.5. J oint compression force Pcomp = 0.5P work —5000 kgf Total stud tension force [Eq. (1.22)J Ptens = (1 + 0.5) P worh = 15 000 kgf Tension force per stud n 15 000 l.~t Ptens. l— / —3750 kgf Tensile stress in studs Gtens 3750 0.785-182 = 15 kgf/mm 2 Compressive stress in cylinder block a, comp 5000 comp - F 2 7000 = 0.7 kgf/mm 2 Total tightening force Pfight [Eq. (1.24)] required Ptight = Pwork (o~\-- Factor E 1 F 1 Factor E 2 F 2 Hence, V ■ e z f 2 ) EiFf—21 000 4-0.785-18 2 = 2.1-10 7 kgf E 2 F 2 = 7500 ■ 7000 = 5.3 -10 7 kgf 1 Ptight —i0 000 I 0.5-f 1 + 2.1 5.3 = 12 000 kgf 3 * 36 Chapter 1. Tightened Connections Angle through which nuts must be turned when tightening [Eq. (1.38)] v=360°P, /gW —( 0 25E(fi +1^) = = 360 °- 3000 ~ ( 1 -± w + lp ^ w ) _ ®- Thermal stress in the joint when heated [Eq. (1.25)] Pt = («2*2 - «l*l)- E %\p~ fey hypothesis, t 2 = tj = 80°C. Substituting numerical values, we have ? 1.107 P t = 80 (22 -11) • lO- 6 ■ ■ -f , ■ = 13 000 kgf 14-—. ^ 5.3 After heating: stud tension force p, ten» = Ptens + Pt = 15 000+13 000 = 28 000 kgf joint compression force P'comp = Pcomp+Pf = 5000 +13 000 == 18 000 kgf tensile stress in studs P’ a' ten , = _ tem ■ te 4-0.785d 2 4-0.785-182 = 28 kgf/mm 2 compressive stress in cylinder block <^mp=^T=2.6 kgf/mm 2 tension force pulsation amplitude A ten, = P'tens ~ ( P tight +Pt) = 28 000 - (12 000 + 13 000) = 3000 kgf coefficient of cycle skewness Pfenj A tens 3000 . „ r ~~ JZZ compression force pulsation amplitude A C omp = Plight +Pt - P’ comp = 12 000+13 000—18 000 = 7000 kgf coefficient of cycle skewness Pcomp 18 000 _ ~~ PfigAf+Pf - 12000+13000 _U ’ 7 ' The fatigue limit of steel grade 30XTC under a pulsating cycle with a coef¬ ficient of skewness of 0.9 equals 75 kgf/mm 2 . The fatigue limit of aluminium alloy grade AJI5 under a pulsating cycle with a coefficient of skewness of 0.7 equals 8 kgf/mm 2 . Consequently, the safety factor for the studs is — = 2.7 and that 40 g for the cylinder blocks, ="3=3. 4.0 Chapter 2 Press-Fitted Connections Press (drive) fits are extensively applied in general engineering to make permanent joints or those seldom dismantled. The relative displacement of press-fitted parts is prevented by the forces due to elastic compressive deformation (in the male part) and tensile deformation (in the female part), which are proportional to tlie amount of interference in the joint. 2.1. Press (drive) Fits The USSR State Standard TOCT 7713-62 stipulates the following press fits for the basic hole system. 1st grade of accuracy 2nd grade of accuracy 1st heavy drive fit (DM i) . 2nd heavy drive fit (Dh2 x ). .. n P u .. np2 t Light drive fit (Dl) .... Heavy drive fit (Dh) . . . Shrink fit ( Sh) . . Ha . lip • r P 2a grade of accuracy 3rd grade of accuracy 1st heavy drive fit (Dhl 2a ) • ■ 2nd heavy drive fit ( Dh2 ia ).. • n Pha • Hp2 2a 1st heavy drive fit (Dhl 3 ) .. 2nd heavy drive fit ( Dh2 3 ) ., 3rd heavy drive fit ( Dh3 3 ) ., ■ n P i 3 ,. Hp2 9 .. Up $3 Figure 16a gives mean interferences A mean pm as a function of the shaft diameter d mm for different press fits; Fig. 166—mean relative interferences pm/mm and . From Fig. 166 it is seen that the relative interferences sharply increase in the small diameter range. For this reason one must be especially careful when designing small diameter joints since the strength of press-fitted parts depends first of all on the amount of their relative interference. Light drive fits ( Dl) are obtained with the least interferences. Fits Dhl 1 are close to the latter in terms of the mean interferences. Fits Dh2 x , Dh, Dhl ia and Dhl z are practically the same as far as 38 Chapter 2. Press-Fitted, Connections the mean interferences are concerned, the only difference being that the fits to lower grades of accuracy have wider tolerances. Mean interferences are practically the same in the case of fits Sh and Mtn Fig. 16. Mean interference S mean and relative interference A moan ld as a func¬ tion of connection diameter d for different fits Dh2 3 . The greatest (and almost identical) mean interferences are in the case of fits Dh2 2a and Dh3 :i . The relationship between the mean interference and diameter is approxi¬ mately expressed by the formula Ajnean pm « 1 1' mm + 60) (2.1) The proportionality factor ip for various fits is as follows Fits. Dl Dhli Dh Dh2i Dhl 3 Dhl 2a Sh Dh2 3 Dh,2 2a Dh3 3 f. 0.23 0.28 0.46 0.44 0.5 0.54 0.95 0.98 1.13 1.16 2.2. Strength of Press-Fitted Connections The axial holding power of a press-fitted connection is Pax = kFf kgf (2.2) The torsional holding power Afior. = 0.001*F/4 kgf.m (2.3) where k — unit pressure on the surface of joint, kgf/mm 2 F = jt dl = joint surface area, mm 2 2.2. Strength of Press-Fitted Connections 39 / = coefficient of friction between mating surfaces (for steel and cast iron, on the average, / = 0.1-0.15) d, l = diameter and length of the surface of joint, mm The unit pressure on the surface of joint is found from Lame’s equation _1_ c l — Pi i C 2+P2 Ei ' El kgf/mm 2 (2.4) where 4- = relative diametral interference (4- = , 77 ,4 ^ ^ d \d 1000 d mm/ Ei,E 2 and p lf p 2 = elastic moduli (in kgf/mm 2 ) and Poisson’s ratios of the materials of the male and female parts, respectively c x and c 2 = coefficients expressed as Cl (2.5) Cl ■+(ir '-nr ( 2 . 6 ) where d x and d 2 are the inner diameter of the male part and the outer diameter of the female part, respec¬ tively. Let us designate = a x and — a 2 . The quantities a x and a 2 may be called the relative wall thickness of the male and female parts, respectively. When a x = a 2 = 1 the thicknesses of the male and female parts equal zero. Values a x = a 2 = 0 correspond to the case of massive male and female parts. Coefficients c x and c 2 may be expressed in a general form as follows (Fig. 17): Equation (2.4) shows that unit pressure k and, consequently, the joint strength are proportional to the relative diametral inter¬ ference A Id, increasing with the increase of Young’s moduli of the materials and decreasing with the increase of c x and c 2 , i.e., with the enhancement of the wall thickness factor a. 40 Chapter 2. Press-Fitted Connections The compressive stress in the male part is maximum on its internal surface and is expressed as 2k A 2 1 a i ' d i—a\ ci — pi c 2 + n 2 Ei "T E 2 (2.7) The maximum permissible unit pressure determined by the bearing strength of the material is k = Ob e ar ( 2 - 8 ) where a bea r is the ultimate bearing strength of the material. The reduction in the internal diameter of the male part (2.9) The tensile stress in the female part is maximum on its inter¬ nal surface and is expressed as 2k A 2 a 2 — 1 _ a *~ — *1 X d 1 — a\ 1 m Ei ( 2 . 10 ) c 2~fP2 Ez The increase in the external diameter of the female part ( 2 . 11 ) Fig. 17. Coefficient c as a function of wall thickness factor a Equations (2.4), (2.7) and (2.10) enable some general con¬ clusions to be drawn. , To simplify the problem as¬ sume that both the [female and male parts are made of the same material (E i — E 2 = E; = = p 2 = p). Then Eqs. (2.4), (2.7) and (2.10) become , A E K ~ d ' Cl + c 2 kgf/mm 2 ° l = ~d' A d ’ E 1 — a\ Oo = — c l + c 2 E C1+C2 kgf/mm 2 kgf/mm 2 ( 2 . 12 ) (2.13) (2.14) Figure 18a shows as a function of a x and a 2 the relative pressure k 0 = —-— which is the value of pressure k when AEld = 1. c i + c 2 2.2. Strength of Press-Fitted Connections 41 The pressure (and consequently the strength of the joint) is maximum when % = = 0 (both the male and female parts are massive) and decreases with the increase of a x and a 2 (i.e., with the decrease of the wall thickness of both the male and female parts), tending to zero when fli = a 2 = 1. The reduction of the pressure with the decrease of the wall thick¬ ness of the male and female parts can be compensated for by increas¬ ing the diameter and length of the surface of joint. If, as is usual. Fig. 18. Effect of wall thickness factors a, and a 2 on values of k 0 , a 01 and a 02 . the length of connection is proportional to its diameter, i.e., I = = nd (n = coefficient of proportionality), then, according to Eqs. (2.2) and (2.3), P ax = kf red 2 and Mtors = kfn Hence, the resistance of a joint to an axial shift is proportional to the square of its diameter, and its resistance to torsion, to the cube of the diameter. From Eqs. (2.13) and (2.14) the relative stresses o 01 and o 02 (the stresses with A Eld = 1) are 1) and increase as the wall thickness of the male part is reduced (aj -> 1); 42 Chapter 2. Press-Fitted Connections stresses cr 02 in the female part (light lines) are maximum (o 02 = 1) with a massive male part (a x = 0); they decrease as the wall thick¬ ness of the male part is reduced (a x —1) and increase as the wall thickness of the female part is reduced (a 2 -> 1). Assuming that the male part is a shaft, and the female part, a hub, the following rules can be formulated: to increase the strength of the shaft it is necessary to increase the thickness of its walls and decrease the wall thickness of the hub (massive shaft and thin-walled hub). This rule is applied when the hub strength is of no concern; to increase the hub strength it is necessary to increase the thick¬ ness of its walls and to decrease the wall thickness of the shaft (mas¬ sive hub and thin-walled shaft). The rule is applied when the shaft is of sufficient strength. When the male and female components of a joint are made from dissimilar materials, the above relationships are affected by the rigidity of the materials. If one of the parts (female or male) is made from a material having a smaller elastic modulus ( E ') than the other ■one ( E "), then the unit pressure and the stresses in the parts fall approximately in proportion to the ratio E'/E” (to a greater degree in the part made from the material of the greater elastic modulus). The best relationship between the wall thicknesses of the hub and shaft should be established in each particular case by individual ■calculations. (a) Coefficient of Friction The strength of press-fitted connections is directly proportional to the coefficient of friction between the mating surfaces [see Eqs. (2.2) and (2.3)]. The coefficient of friction depends on the pressure on the contacting surfaces, the magnitude and profile of surface microirregularities, the material and state of the mating surfaces (presence of oil) and also on the assembly technique (connection with the aid of a press, with heating or cooling the parts). The coefficient of friction rises with an increase in the surface roughness and lowers with a rise in the unit pressure. It is higher when the assembly is done with heating or cooling as compared with a press-fitting. The coefficient can substantially be enhanced by -electroplating. Depending on the above factors the coefficient of friction ranges from 0.08 to 0.3, and sometimes may be even higher. Obviously no accuracy in calculations can be expected with such an extensive / value scatter. The primary value of the calculation is that it enables one to determine the influence of the geometry and rigidity of the •connection elements on the joint strength and outline rational ways to improve it. 2.2. Strength of Press-Fitted Connections 43 In practice lower coefficients of friction (/ = 0.1-0.15) are gene¬ rally adopted, including possible increase in the coefficient above these values- in the safety factor. (i b ) Effect of Surface Finish An adequate surface finish of the mating components is an essential condition for the strength of pressed connections. The measured hole and shaft diameters include the height of microirregularities which are flattened in the course of press-fitting. i_i_i_i_l_i_l, ...J_l_l_l 0 10 20 30 60 50 60 70 80 d,mw Fig. 19. Mean interferences with Dh grade fit (heavy line) and roughness heights for various classes of surface finish as a function of connection diameter d If the height of surface microirregularities is commensurate with the interference of mating parts, the actual interference in the pressed connection will then be appreciably smaller. Shown in Fig. 19 are the mean values of interferences in the case of a heavy drive fit (Dh) for different shaft diameters, and the total roughness height 2 Rz of the shaft and the hole when machined to the 4th-8th class of surface finish. From the graph it is clear that small-diameter connections (less than 30-40 mm) must not be machined worse than to the 6th class because the total roughness height in this case becomes close to the interference value. Hence, in such connections the interference will either be substantially reduced or disappear altogether after the microirregularities are flattened. 44 Chapter 2. Press-Fitted Connections Connections with diameters of the order of 50 mm and more, as well as connections with greater interferences, may be machined somewhat rougher. In practice the surfaces of the mating parts in medium-size press-fitted connections are machined to the 8th-10th and 7th-9th classes of surface finish (for shafts and holes, respec¬ tively). Microirregularities have a certain positive effect on the strength of the connection as they act like spurs enhancing the bond between the contacting surfaces. Experiments show that machining better than to the 11th class of surface finish impairs the joint strength because of the reduced coefficient of friction between the contacting surfaces. Design Eqs. (2.4), (2.7) and (2.10) include the actual interference value. Therefore, when designing pressed connections the assigned nominal interference A nom should be reduced by the amount of the expected flattening of microirregularities A* = 2

ecomes acceptable: o 2 = 0.5-10.6 = 5.3 kgf/mm 2 . Let us now consider the case of pressing a disk of a wrought aluminium alloy onto a hollow steel shaft with external diameter d = 100 mm and internal -diameter d x = 70 mm (% = 0.7). The disk can be regarded as a solid part ia 2 = 0). The grade of fit is Dh (A = 65 pm). The class of surface finish of the -shaft is the 9th ( R zl = 1.6 pm), and that of the disk bore, 8th ( R z2 — 3.2 pm); R n + R z2 = 4.8 pm Af?2 d (65—4.8)7200' 1000-100 ■ 4.3 kgf/mm 2 From the diagram we find (point b) that*o = 0.465; o 01 = 1.82; 0 O2 = 0.92. Hence * = 0.465-4.3 = 2 kgf/mm 2 01 = 1.82:4.3=8 kgf/mm 2 02 = 0.92-4.3 = 4 kgf/mm 2 Fig. 22. Diagram for computing press-fitted joints (steel parts pressed in parts of aluminium alloys) 54 Chapter 2. Press-Fitted Connections Assume that during operation the disk temperature rises by 100°C (in compari¬ son to the assembly temperature) and the shaft remains cold. As the linear expansion coefficient of an aluminium alloy a 2 = 22-10~ 6 , the disk bore dia¬ meter will increase in heating by the value A< = a2td‘10 3 r22-100-100-10 3 106 = 220 pm Thus, the original press-fit interference vanishes and instead a clearance amounting to 220—(65—4.8) = 160 pm develops in the joint. To keep the alignment of the parts a fit of greater interference should be employed, for example, that of the grade Dh2 2a (A = 180 pm). In this case during heating of the joint there develops a clearance amounting to 220—(180—4.8) = = 45 pm which does not disturb the alignment. With an interference of 180 pm d (180 - 4.8) 7200 1000-100 = 12.6 kgf/mm 2 The values of k, o 1 and o 2 increase by a factor of ^ /y ss 3. Stress o 2 in the ■disk hub (in cold state) becomes o 2 = 3-4 = 12 kgf/mm 2 , which is acceptable for wrought aluminium alloys. Pressing bronze components into steel ones (Fig. 23). A tin-bronze bush with external diameter d = 40 mm and internal diameter d t = 35 mm (a t = 0.87), is press-fitted into a steel hub with external diameter d 2 = 53 mm (a 2 — 0.75). The grade of fit is Dh2 2a (A = 80 pm). The joint surface of the bush is machined to the 9th class of surface finish (i? zl = 1.6 pm) and that of the hub, to the 8th class (H z2 = 3.2 pm); fi zt + R z2 = 4.8 pm A E z (80 - 4.8)21 000 d ~ 1000-40 40 kgf/mm 2 From the diagram we find that when a r = 0.87 and a 2 = 0.75 (point a), * 0 — 0.06; o 01 = 0.5; o 02 = 0.27. Hence * = 0.06-40 = 2.4 kgf/mm 2 = 0.5-40 = 20 kgf/mm 2 a 2 = 0.27-40 = 10.8 kgf/mm 2 Stress a, in the bush exceeds the yield limit of tin bronze in compression (o 02 = 15 kgf/mm 2 ). We reduce the internal diameter of the bush to d x = 30 mm (a 2 = 0.75). On the diagram (point b) we find that *„ = 0.1; o 01 = o 02 = 0.46. Hence * = 0.1-40 = 4 kgf/mm 2 Oj = o 2 = 0.46-40= 18.4 kgf/mm 2 It is clear that the increase of the bush wall thickness is of little effect; the stresses decrease only by 8% and still exceed the yield limit of the material. The reduction of the hub wall thickness does not help either. Let a 2 = 0.85 (d 2 = 47 mm). From the diagram o 01 = 0.35, whence o x = 0.35-40 = - 14 kgf/mm 2 . Let us now try to reduce the interference. We take the Dhl 2a grade fit 50_^ g (A=50 pm). Then, the actual interference decreases by a factor of =— -2— = 0.6; oU-—4.0 the stress in the bush (with the original a t = 0.87) then becomes acceptable: 1 56 Chapter 2. Press-Fitted Connections Ox = 0.6-20 — 12 kgf/mm 2 , and when a 2 = 0.85 it becomes a, = 0.6-14 = = 8.4 kgf/'mm 2 . Now assume that during operation the joint temperature rises by 100°C. The coefficient of linear expansion of bronze is = 18-10“ 8 ; that of steel, a 2 = 11-10 -6 . The temperature interference will then be A t = 1000-100-40(18 —ll)-10- 6 =28 pm The interference in the joint A = 50—4.8 -f 28 = 73 pm. 73 The unit pressure and stresses are increased by ^—— = 1.6 times. For ou—4.o the bush with a 1 = 0.87 the stress becomes aj = 1.6-12 = 19 kgf/'mm 2 and still exeeds the yield limit of the material. Now we take the Dh grade fit (A = 40 um). In contrast to the previous case, 40 — 4 8 the actual interference decreases by a factor of —~ 0.49 and the stresses in the bush reach an acceptable level: a 2 = 0.49-19 = 9.3 kgf/mm 2 . Pressing bronze components into those of cast iron (Fig. 24). A bronze bush having the same parameters as in the previous example (d = 40 mm, a x = 0.87) is pressed into a cast-iron hub (a 2 = 0.75) by the Z)h2 2a grade fit (A = 80 pm). The surface finish is the same as previously (R tl + R z n = 4.8 pm) __ ---= 15 kgf/mm 2 On the diagram we find that for a 2 = 0.87 and a 2 = 0.75 (point a), k 0 = = 0.108; Ooi — 0-91 and 002 = 0.5. Hence k — 0.11 -15 = 1.65 kgf/mm 2 a t = 0.91-15 = 13.6 kgf/mm 2 a 2 = 0.5-15 = 7.5 kgf/mm 2 Owing to the lower modulus of elasticity of cast iron the stresses in this case are fairly lower than those encountered when pressing the bush into a steel com¬ ponent (the previous example). Nevertheless the stress in the bush is close to the yield limit of bronze. Let the grade of fit be Dhl 2a (A= 50’pm). The actual inter¬ ference then decreases by a factor of •=:—= 0.6 and the stress in the bush becomes 04 = 0.6-13.6 - 8.2 kgf/mm 2 . Let the joint temperature rise in operation by 100°C. In the joint there develops a temperature interference equal as previously to 28 pm (the coefficient of linear expansion of cast iron is approximately the same as that of steel). According to the previous example, the stress in the bush will increase by 1.6 times and become Ox = 1.6 -8.2 = 13 kgf/mm 2 (in comparison with ct x = = 19 kgf/mm 2 when heating a steel hub). However, in this case too it is recom¬ mended that the interference be reduced further. Now we use the Dh grade of fit. Then, according to the previous conditions, the stress will drop by a factor of 0.49 and become = 0.49-13 = 6.4 kgf/'mm 2 (in comparison with Ox = = 9.3 kgf/mm 2 for a steel hub). Now consider the case of a solid cast iron part (a 2 = 0). Assume that the bush parameters remain as before (a 2 = 0.87). The grade of fit is Dh (A = 40 pm). On the diagram (point b) we find that k 0 = 0.15; cr 01 = 1.25 and 0 O2 = 0.3. A E 2 (40 — 4.8) 8000 d 1000-40 7 kgf/mm 2 0.05 OJ 0.15 01 015 0.3 0.35 01 015 0.5 0.55 0.6 0.65 Fig. 24. Diagram for computing press-fitted joints (bronze parts pressed in cast-iron parts) 2.5. Probability Calculations for Press-Fitted Connections 59 Hence k = 0.15-7= 1.05 kgf/mm 2 =1 - 7 = ( a^f 2 ji J e 2• 0.5 the risk percentage is negligible. Thus, when Z = 0.7, for every 1000 connections approximately one is a risk, and when Z = 0.6, approximately five connections have parameters exceeding the given limits. Consequently, there is very little risk in narrowing the margin tolerance to Z8 and introducing in place of the extreme nominal deviations A mln and A max probable deviations Amin — A min" 6(1—Z) and Amai —— An 6(1 -Z) where Z is a value ranging from 0.9 to 0.5, depending on the adopted risk percentage. Let us carry out a comparative numerical computation of a press-fitted connection by both conventional and probability methods. Assume a connec¬ tion comprising a solid steel shaft, 80 mm in diameter, and a steel sleeve with an external diameter of 120 mm. The connection length Z = 80 mm. The grade of fit A/Dh. The shaft and hole are machined to the 8th class of surface finish (R z = 3.2 pm). Hole tolerance:+30 urn; shaft tolerance: lower limit +45 pm, upper limit +65 pm. Coefficient of friction / =0 .1. Nominal interferences: maximum 65—0 - 65 pm, minimum 45—30 = 15 pm. With a microirregularity flattening coefficient of 0.5, the flattening amounts 62 Chapter 2. Press-Fitted Connections to 6.4 (xm. The actual maximum interference is, thus, 65—6.4 = 58.6 pm, and the minimum, 15—6.4 = 8.6 pm. For the probability computation assume that Z = 0.6 (risk percentage \—Z is 0.517). The value -equals 0.2. Probable deviations of dimensions: hole — minimum 0 + 30-0.2 = 6 pm, maximum 30 — 30-0.2 = 24 pm, shaft—maxi¬ mum 65 — 20-0.2 = 61 pm, minimum 45+ 20-0.2 = 49 pm. Probable inter- erences: maximum 61 — 6 =55 pm, minimum 49 — 24 = 25 pm. Accounting for fthe microirregularity flattening the maximum interference is 55 — 6.4 = = 48.6 pm, and the minimum interference, 25 — 6.4 = 18.6 pm (Table 5). As seen from Table 5 the probability computation gives more favourable characteristics which at the same time are closer to the actual parameters of completed connections. The normal distribution law is valid wherever a large number of occurrences is involved, and, hence, for mass production and with controlled operations at that. Substantial deviations from this law occur in individual and small-lot production, attributable in the first instance to the small number of occurrences and, secondly, to certain specific features of the machining process. With individual production the operator involuntarily keeps to the lower hole toler- 2.5. Probability Calculations for Press-Fitted Connections 63 Table 5 Calculation Results for a Pressed Connection Characteristics Computation by nominal values probable values Maximum interference, pm 58.6 48.6 Minimum interference, pm 8.6 18.6 Press-fitting force (with maximum interference), kgf Holding power (with minimum interference): 8500 7200 axial, P ax , kgf 1250 2750 Aftorsi kgf-m Stresses (with maximum interfe¬ rence): 50 110 in shaft, a u kgf/mm 2 8.6 7.1 in hub, a 2 , kgf/mm 2 15.3 12.7 ance limit and the upper limit for the shaft, orienting to the NO-GO side of gauges. Because of this the holes are made, on the average, closer to the minimum (nominal) size, while the shaft dimensions show a tendency toward the ma¬ ximum (upper) limit of toleran¬ ce. The grouping centres on the distribution curves are thereby shifted (Fig. 28) and the proba¬ bility of obtaining maximum interferences is thus increased. An asymmetry of size distri¬ bution also periodically occurs in mass production with cont¬ rolled operations. As the cutting tools wear down, hole dimen¬ sions approach the minimum, and shaft dimensions, the maximum. Fig 2 8. Distribution curves with The periodicity of this pheno- displaced, grouping centre menon depends on operation ad¬ justment frequency and is absent only with automatic readjust¬ ment. To establish any regularities of the dispersion changes in a general form is difficult. However it is still necessary to note that any probability com¬ putation gives average figures of the expected size distribution over a long time interval and for large batches of products. The possibility 64 Chapter 2. Press-Fitted Connections of a temporary crowding of low-probability size combinations is not excluded, as a result of which comparatively large batches of defective connections will appear, although the mean level of risk related to very large batches will still remain within the computed limits. All this restricts the practical value of the probability computa¬ tion technique. When designing press-fitted connections it is better to keep within narrow boundaries such an interference as will ensure good working capacity of the connection without inducing high stresses in the mating components. With the existing systems of press fits, which are noted for their large deviation scatter, it is difficult to achieve. The selective assembly method is undesirable as it complicates production. The general rule is to employ fits to the highest grades of accuracy, in particular, those to the first grade of accuracy. However, this class of fits only covers a limited range of interferen¬ ces. Therefore it seems advisable to develop a single class of press fits with lower margin tolerances that would encompass the entire range of interferences necessary for the industry. 2.6. Press-Fitting with Heating or Cooling of Parts The press-fitting force may reach substantial values, especially with great interferences and large surfaces of joint. This force pro¬ gressively increases as the part being press-fitted is forced deeper into the hole, reaching its maximum at the end of the procedure. The maximum press-fitting force may be determined from Eq. (2.2). Let us find the force necessary to press a solid steel shaft (a x = 0) with diameter d = 100 mm into a cast-iron hub 150 mm long and having external diameter d 2 = 165 mm (a 2 = 0.6) by the Dhl ia grade fit (A = 90 pm). On the diagram shown in Fig. 21 we find that for a x = 0 and a 2 = 0.6, k 0 = 0.39. The unit pressure k = k 0 A E 2 d = 0.39 90-8000 1000-100 = 2.8 kgf/mm 2 The maximum press-fitting force is P = kfF = 2.8 • O.Iji • 100 • 150 = 13000 kgf The press-fitting procedure can be facilitated by heating the female part or cooling the male part, or else by practising both of these at once. When press-fitting parts into large housing structures only the male component is cooled. The heating temperature for the female part to have zero inter¬ ference A 1 2.7. Pressed Connections with Electrodeposited Coatings 65 where t and t 0 = heating temperature and workshop temperature, respectively ^= relative interference in the joint cc 2 = linear expansion coefficient of the female part material d = diameter of the joint, mm Similarly, when cooling the male component i '+^ = 4"a 1 1 1000 (2,25) where t' is the cooling temperature. It should be remembered that at subzero temperatures the linear expansion coefficient a decreases (see Fundamentals of Machine Design, Vol. 1, Fig. 257). Assume that in Eq. (2.1) A mean «ij)D and that f 0 = 20°C. Then, from Eqs. (2.24) and (2.25) one can find temperatures t (heating) and t' (cooling) required to obtain zero interference when assembling joints by different fits. For cast-iron and steel components (a « « 10 -6 ) these temperatures are as follows: Fits .... Dhli Dh and Dh2^ Dhl 3 Dhl Za Sh Dh2 3 Dh2 2a Dh3 s t,° C . . . 43 60 65 68 105 108 122 124 t', °C . . . —8 —30 —36 —40 —86 —90 —105 —110 Generally the cooling procedure is effected by using dry ice (evap¬ orates at — 80°C); or, for more intensive cooling, liquid nitrogen (—196 °Ci and oxygen (—183°C). It must be borne in mind that heated components rapidly cool as they are being delivered from the furnace to the press. Furthermore, during the press¬ fitting operation the temperature of the heated hub quickly lowers due to its contact with the cold shaft. Therefore, due corrections must be introduced to compensate for the decrease of temperature during handling and press-fitting. In practice this implies that the components to be press-fitted must be overheat¬ ed by 30-50°C as compared with the calculated temperature. The cooling temperature should also allow for the heating effect. 2.7. Pressed Connections with Electrodeposited Coatings The strength of press-fitted connections (axial and torsional holding power) can significantly be increased by electroplating the contacting surfaces. Figure 29 shows the results of comparative tests of pressed connections assembled by the A/Dh grade fit. The electrodeposited layer on the contacting surfaces was 0.01-0.02 mm thick. Two assembly procedures were used: one by means of a hyd¬ raulic press and the other with cooling the shaft in liquid nitrogen. In the latter case the radial clearance between the mating surfaces was 0.05 mm. 5—01418 66 Chapter 2. Press-Fitted Connections A connection without coating assembled in a hydraulic press (without shaft cooling) was used as a reference specimen, its axial holding power P 0 being taken as the unit of comparison. From the diagram it is seen that electroplating sharply (by 2-4.5 times) enhances the connection strength, and that the assembly with cooling the shaft assures a much higher connection strength than f 0.8 0.7 0.6 0.5 OA 0.3 0.2 0.1 Without Cd Cu In NL Cr coating Fig. 29. Relative strength of pressed connections with deposited coatings. Blackened columns indicate assembly with shaft cooling, while nachured columns are for assembly in a press. The strength of a pressed connection without coating and assembled in a press is taken as unity (after G. A. Bobrovnikov) the press-fitting assembly. For uncoated connections assembled with cooling the joint strength is doubled, and for connections with soft coatings (Cd, Cu, Zn), improved by 20-30% in comparison with the assembly in a press. The strength of connections with hard coatings (Ni, Cr) and assembled with cooling is worse than that of the same connections assembled in a press. The greater strength of connections with electroplated coatings is probably due to the occurrence of strong molecular bonds at the joint surfaces. The pro¬ longed dwell under the high pressure existing on the contacting surfaces causes a diffusion process whereby the atoms of the base and coating metals mutually penetrate each other, thus forming some intermediate structure. In other words, a “cold welding” process seems to occur. As for the rather high (approaching unity) values of the coefficient of friction (see the right-hand ordinate of the diagram in Fig. 29), this can readily be explained. The friction coefficient concept in its classical mechanical interpretation in these conditions is no longer valid: the value of the friction coefficient here shows the shear resistance 2.8. Design of Pressed Connections 67 of the intermediate layer associated with the base and plated metals and not, as is usual, the resistance to the displacement of one surface relative to another. The lower strength of connections assembled with the aid of a press is attri¬ buted to the fact that the ridges of surface microirregularities are flattened and/or cut away. During the assembly with cooling the ridges remain intact and after heating penetrate the valleys in the mating surface, thus adding to the connection strength. Joints with soft electroplating are disassembled without damage to the surfaces, but when disassembling connections with hard coatings, scores, cracks and deep tears in the base metal are observed, sometimes over considerable areas of the contracting surfaces, due to which reassembly is difficult or even becomes impossible. The situation is aggravated by the fact that hard electroplating lowers the fatigue strength of the joint. Due to all these reasons preference should be given to soft coatings. The application of soft coatings in conjunction with the assembly with cooling allows the connection strength to be improved 3-4 times as compared with that of the press-fitted connections without coat¬ ings. Thus, it becomes possible, while maintaining the given joint strength, to employ smaller interferences, reducing, in this way, the tensile stresses in the female part and compressive stresses in the male part. In addition, the electrodeposited coatings protect the fitted surfaces against corrosion and eliminate the “welding” of surfaces subjected to cyclic loads. The strength of pressed connections can seemingly be improved by metalli¬ zation and thermodiffusion saturation (e.g., by thermodiffusion zinc plating) of fitting surfaces. Still further strengthening of pressed connections can probably be attained by depositing dissimilar coatings upon the joint surfaces, for instance, a zinc coating on one surface and a copper one on the other. From the atomic diffusion of the two metals so.ne intermediate structures can be expected in the contact zone, possessing higher strength than homogeneous platings (e.g., brass-type alloys when zinc and copper deposits are combined). 2.8. Design of Pressed Connections The specific feature of pressed connections is that their joint surfaces are already prestressed by interference forces prior to the application of working loads, subjecting the female part to a biaxial tension which is unfavourable from the strength viewpoint. When adding the prestresses to the working stresses, their resultant may exceed the yield limit of the material and cause connection failure. Moreover, formal calculations of press-fitted joints, based on the assumption of constant cross-sections throughout the entire part length where the end conditions can be ignored does not express true stress values. In practice the strength of a joint strongly depends on the shape of the female and male parts. Uneven rigidity of parts (stepped shafts, hubs with disks, etc.) causes an unequal distribution 68 Chapter 2. Press-Fitted. Connections of contact pressures and stresses over the connection’s length. Sharp stress gradients arise at the joint edges. Formal calculation, even with high safety margins does not always assure a reliable connection. The more so if one takes into account that the magnitude and distribution of working stresses throughout the part cross-section and their relationships with prestresses remain mostly uncertain, especially in joints subjected to cyclic stresses. Therefore, regardless of computation results all constructive measures to strengthen pressed connections must be taken. In order to enhance the strength and reliability of pressed con¬ nections it is advisable to: enlarge the diameter and increase length of connections so that unit pressure upon contacting surfaces is reduced; < a > lb) (c) id) ie> If) Fig. 30. Lead-in chamfers and recesses in pressed connections select interference values within narrow magrins, making use of more accurate fits; avoid sharp changes in cross-sections of the connected parts over their contact areas (areas close to these parts), thus preventing abrupt stress relations; improve contact surfaces by appropriate heat treatment (e.g., low temper hardening, induction hardening, etc.) or strengthen them by mechanical hardening (shot blasting, rolling of shafts, rolling out or burnishing of holes); use assembly technique in which the female part is heated or the male part is cooled; electroplate contacting surfaces, using soft metals (Zn, Cu, Cd). The reliability of pressed connections depends greatly on correct assembly. To facilitate the press-fitting procedure leading chamfers should be provided on the shaft and in the hole. The chamfers should be made at 45° (Fig. 30a) or, better, at 20-30° (Fig. 306, c). If great interferences are expected, the shaft must have more gradual cham¬ fers of 5-10° (Fig. 30d). The chamfer leading diameter d is made 0.1-0.2 mm less than the hole diameter d 0 . Occasionally the shaft or the hole is designed with leading cylin¬ drical collars having an aligning fit, e.g., a slide fit (Fig. 30e, f). 2.8. Design of Pressed Connections 69 The location of the aligning collar in the hole (Fig. 30 f) necessitates the employment of the basic shaft system. It is very important to prevent seizure anti distortion of the joined parts, which impede the press-fitting process or may even spoil the connection altogether. Bush-type thin-walled components are guided when being press- fitted by means of a centring mandrel (Fig. 31a). Figure 31b shows Fig. 31. Methods for press-fitting thin-walled bushes Fig. 32. Methods for axial fixing of parts when press-fitting a method of press-fitting used for through holes. A bush is mounted upon a screwed-on mandrel with a pilot shank 1 , introduced into the hole with a slide fit. After the press-fitting is over, the mandrel is screwed off. Axial position of the part is fixed by pressing the part home to a stop shoulder (Fig. 32a, b), a hole step (Fig. 32c), flush with the hole edge (Fig. 32c?), or until the hole edge is flush with a shoulder on the part (Fig. 32e). Plain parts can be fixed in any position by means of distance rings 1 placed under the press plunger (Fig. 32/). A typical error made when designing pressed connections is the insufficient length of the joint collar. The connection with a short collar quickly fails due to the crushing of the contact surfaces under the effect of working loads. Examples of wrong and correctly designed pressed connections are illustrated in Fig. 33. For general purposes and rough determination of the minimum length of pressed connections the following formula may be used: ^i„ = 4d 2/3 (2.26) 70 Chapter 2. Press-Fitted, Connections where Z m in = length of the press-fitted portion (less chamfers), mm d = connection diameter, mm A graph based on Eq. (2.26) is plotted in Fig. 34. If the connection is subjected to high bending or shearing loads, especially alternating ones, or if it is necessary to assure a precise Fig. 33. Pressed connections: (a) connecting rod endl (b) press-fitting cup-shaped component on shaft; 1 — wrong; 2 — correct positioning and sound anchoring of the press-fitted component (e.g., bed columns), the length of the press-fitted portion is con¬ siderably increased (l — 1.5-2 d). Avoid, whenever possible, press-fitting in blind holes, because the latter are more difficult to machine and hamper dismantling. Should the fitting in blind holes be unavoidable, it is necessary to assure free exit of air during the press-fitting process. Compression of air when pressing, accompanied by increase of its specific volume, can break the female part, particularly when it is thin-walled or made of a less strong material (e.g., light alloys). Air is vented via holes (Fig. 35a, c) or grooves (Fig. 355). Never press-fit components through two sections with holes of the same diameter (Fig. 36a): while passing the component through the first section (the upper in the Figure) a misalignment frequently occurs, that impedes the introduction of the male part into the second section. In addition, scores may form on the surfaces of the mating parts. Under such circumstances each section should have holes of dif¬ ferent diameters (Fig. 365). The axial dimensions of the connection must be so designed that the part enters the second section first to a depth m = 2-3 mm (Fig. 36c), thus obtaining a definite direction, and only after that enters the first section. It would be wrong to force a bush into a casing as illustrated in Fig. 36 d. Here, in order to reduce the amount of precision machining, 2.8. Design of Pressed Connections 71 the bore has two short joint collars. The error is that both collars have the same diameter. Besides, the distortion of the bush on the fitting areas is inevitable. Imin, mm Fig. 34. Minimum length l m i n of pressed connections as a function of joint diameter d If strict linearity of the hole walls is important, the bush then must be reamed after the assembly, or fitted over its complete length Fig. 36e) or, at least, over a major portion of its length (Fig. 36/). Fig. 35. Air evacuation when press-fitting parts in blind holes Female components must be given ample rigidity to avoid their deformation during the forcing procedure. In the case shown in Fig. 36 g the upper eye will deflect, thereby making impossible the press-fitting into the lower eye. If for some 72 Chapter 2. Press-Fitted. Connections s cannot be made th employ some stiffe: The simplest device etween the eyes. (d) (e) (f) (g) Fig. 36. Press-fitted connections The employment of such a spacer must be thought of beforehand by designing the distance between the eyes to suit one common spacer which can be used for a series of parts. The female and male parts should be of uniform rigidity in the radial direction. Local weakenings, cut-outs, etc., should be avoided. Fig. 37. Press-fitting a bush In the case depicted in Fig. 37 a press-fitting is impeded by the una¬ voidable deflection of the bush toward the cut-out. Furthermore, at the cut-out area the bush will be given additional deflection due to the unilateral radial interference. The situation can be improved 2.8. Design of Pressed Connections 73 slightly if the bush is press-fitted into two sections located in the non-cut-out areas of the hub (Fig. 376). In this case it is better to assemble the bush by a location fit and fasten it in position with bolts (Fig. 37c). Press-fitting is inapplicable when the female or male part has through-cuts extending to the end face (Fig. 37 d). If cut-outs cannot be eliminated the only reason¬ able way is to employ a location fit. Occasionally it is necessary to position press-fitted parts at a certain angular attitude. Such, for example, is the case when press-fitting a keyed shaft into its hub. The alignment of the key with its slot is assured if the leading end of the shaft (Fig. 38a) has a location or free fit over length l which exceeds the dis¬ tance h of the key from the shaft end face. The key is first intro¬ duced into the keyslot and after that the shaft is press-fitted. Another method: the key protrudes out from the shaft to a distance h which is sufficient to assure the shaft location in the keyslot before the press-fitting procedure begins (Fig. 386). The best practice is to assemble the connection with previously heated hub or cooled shaft in order to obtain a clearance in the m Fig. 38. Fitting a keyed shaft in a hub Fig. 39. Press-fitting cams in a disk connection. The angular location of the shaft in the hole will then present no difficulties. Lobed cams with given edge setting angles (Fig. 39) should be pressed in place by means of a guiding adapter having radial cut¬ outs for the cam lobes, which is located from the central hole in the disk. The possibility of applying such an adapter must be considered in designing. 74 Chapter 2. Press-Fitted Connections Figure 39a shows an incorrect design: large shank at the cam root prevents the cam passing through cut-outs in the adapter. In alter¬ native, presented in Fig. 396, the shank size is reduced but the dis¬ tance m between the cam edges is less than the fitting diameter d, therefore, the cut-outs in the adapter must have formed contours. In the correctly designed version, Fig. 39c {m> d) a simple adapter can ensure a correct cam orientation. (a) Pressing-Out jwai Fig. 40. Pressing-out conditions When designing a pressed connections the possibility of its disas¬ sembly must be considered. Parts to be pressed out must have sur¬ faces (preferably flat) which can rest during the pressing- out upon solid plates or bush¬ es. A poor design is illustrated in Fig. 40a. Here, the pulley press-fitted upon the hollow shaft has to rest during the pressing-out procedure on its conically-shaped portion, which complicates the shape of the bearing plate intended to receive the pulley during the pressing-out. The sharp edges of the shaft are unsuitable for resting against the end face of the press ram. In the design shown in Fig. 406 the pulley has a bearing cylindrical web; the shaft end face is made flat. However, in the course of pres¬ sing-out excessive stresses may arise in the pulley disk, particularly if it has a large diameter. Therefore it is better to position bearing surfaces close to the hub (Fig. 40c). The pressing-out force reaches its peak value (rather high) at the initial moment, when the friction of rest is being overcome. After this the pressing-out force decreases because the friction of rest concedes to that of motion, while the length of the press-fitted section continuously decreases as the female part moves off the shaft. Over the last few years hydraulic systems are used in which pres¬ surized oil, whose pressure exceeds the contact pressure (of the order of up to several hundred atmospheres) is admitted into an annular recess in the joint surface through a hole in the shaft (Fig. 41a) or in the hub (Fig. 416). The pressure of oil produces elastic radial deformation of the parts to be pressed out. At the same time the 2.9. Conical Fits 75 presence of oil minimizes friction during pressing-out. Moreover, the layer of oil, penetrating into a circular slot in-between the parts Fig. 41. Pressing-out hydraulically Fig. 42. Pressing-out a bush hydraulically due to capillary forces, acts as a wedge, which greatly facilitates the pressing-out. When pressing-out conical joints hydraulically, no mechanical force is required to force the female part off the shaft. Figure 42 shows how a bush can be pressed-out hydraulically from a blind hole. The bush space is filled up with oil and the bush forced out by a sharp blow on plunger 1. 2.9. Conical Fits Alongside cylindrical press fits extensive use is made of conical press fits (Fig. 43). Tapers K usually range from 1 : 50 to 1 : 100. Fig. 43. Conical pressed connections The required connection interference is obtained by applying a defi¬ nite force when press-fitting the shaft in place. The force is rigorously controlled since the slow taper easily causes excessive radial inter¬ ferences. 76 Chapter 2. Press-Fitted Connections The press-fitting force may be determined from the relation P = Fk (sin a + /) where F = joint surface area, mm 2 k = unit pressure on the contacting surfaces, kgf/mm 2 a — cone generatrix angle (Fig. 43a) / = coefficient of friction between the mating surfaces As the magnitude of sin a is very small in comparison with /, it can be omitted. Then P = Fkf As the values oi D and d are rather close, the joint surface area (i.e., the curved surface of the truncated cone) will be F » nDl Finally P « nDlkf (2.27) Another method is forcing the male component into the female one to a definite depth h (Fig. 43 b), counting from the moment the mating surfaces come into close contact. The amount of h (axial interference) can be found from the relation fe = 10-3 A + 2( P (foi + r z2) = 1Q -3 A+2

) and (d) correct Threaded holes with skew (Fig. 67a) or stepped (Fig. 67c) edges are unacceptable under any circumstances, since it is extremely dif¬ ficult to screw a fastener into such a hole. They can tap such holes only having left beforehand a flat area on the hole end face which will be removed after tapping. 4.4. Reinforcement of Fastening Joints Fastening bolts and studs should be positioned at the rigidity nodes of the assembled parts so that the fastening force is spread over as large an area as possible. Fig. 69. Methods of reinforcing fastening units 7* 100 Chapter 4. Screwed Connections The spacing of the lid fastening bolts shown in Fig. 68a is wrong as the bolts are arranged in bosses positioned in areas weakly con¬ nected with the part body. Somewhat better is the design in Fig. 68 b where the lid has a stiffening edge improving the force distribution over the tightening area. In the correctly designed version (Fig. 68c) the bolts are spaced at the rigidity nodes at the corners. The rigidity Fig. 70. Fastening a bearing cap to crankcase of the lid as a whole is further enhanced by diagonal stiffening ribs, tying up the fastening points with the part body. Figure 68 d shows badly positioned bolts in an intricately shaped flange. The correctly designed version is pictured in Fig. 68e where the bolts are positioned at the rigidity nodal points within the part’s external contour, which enhances clamping rigidity and improves the outer appearance. The bosses in cast parts intended for receiving screwed ends of bolts and studs (Fig. 69) must be reinforced with ribs oriented pre¬ ferably in line with the bolt axis (Fig. 69 b-d). Load-carrying studs and bolts (Fig. 69c), particularly when the clamped parts are made from low-tensile alloys, should be screwed deeper into the parts 4.4. Reinforcement of Fastening Joints 101 so that the maximum cross-sectional area of the wall is engaged in the work. Furthermore, longer studs under cyclic loads are stronger and ensure a more reliable fastening. Figure 70 shows the subsequent stages of strengthening a crank¬ case and bearing cap assembly loaded in tension. The weakest design is that with the short studs (Fig. 70a). Here is a further defect, namely, the crankcase strengthening ribs are offset relative to the stud axes and do not share the load. In improved versions (Fig. 70b, c) the studs are positioned deep into the case body and the bosses reinforced with ribs. In the design illustrated in Fig. 70 d the studs ends are tightened with nuts, relieving the case thread from load; the crankcase cross-section, which works in tension, is reinforced with an arch-shaped strengthening collar. Studs can be relieved from bending (in one plane) by installing their screwed ends in cylindrical swivelling inserts (Fig. 70e). The most advisable design (Fig. 70/) has the stud ends outside. The tensioning of the material inherent in the previously illustrated versions is practically eliminated and the load is taken up by the entire crankcase cross-section. Chapter 5 Flanged Connections When designing flanged connections it is necessary to ensure strength and rigidity with the minimum weight, as well as the rigidity of the sections connecting the flanges to the walls of the part. Illustrated in Fig. 71 are typical flange designs for turned steel cylindrical parts (a sleeve held to a housing), approximately in the order of increasing rigidity. The design in Fig. 71a is unsatisfactory: the flange connection to the walls is too thin and inadequately rigid. The principal flange strengthening methods are: introducing fillets (Fig. 71 6, c) and tapers (Fig. 71 d-f) in the areas where the flanges are connected to the walls. When large fillets and tapers are used, nut bearing surfaces are machine faced (Fig. 71c, e, f) so that the fastening studs may be brought closer to the walls. The weight of flanges is reduced by holes machined between the fastening studs (Fig. 72a), by removing superfluous material off the periphery (Fig. 726, c) and surface (Fig. 72 d,e), and by face recesses (Fig. 72/, g) and radial recesses (Fig. 72 h). Typical designs of cast flanges are pictured in Fig. 73. The rigidity of flanges is improved by ribbing (Fig. 736), by making local bosses in the fastening hole areas (Fig. 73c), by making the flanges higher (Fig. 73 d-f). To eliminate excessive solidness high flanges are under¬ cut. Flanges with through undercuts (Fig. 73 d) have the disadvantage that the flange is subjected to bending when tightening fastening studs. This shortcoming is eliminated in the design having bosses around the fastening holes (Fig. 73e, /). Flanges can be reinforced with a continuous peripheral rib (Fig. 73 g) linked by transverse ribs to the walls of the part. Occasion¬ ally a flange is made in the form of two flanges interconnected by bosses of the fastening holes (Fig. 73 h). Rigidity is significantly enhanced by locating fastening studs in recesses, which have semi-circular cross-sections (Fig. 73 i-k). Still further improvement can be attained by increasing the height of fastening stud bosses (Fig. 73 l, m). In the designs depicted in Fig. 73 n-p the stud-receiving bosses are made in the form of columns; the nut-bearing surfaces are slightly 104 Chapter 5. Flanged Connections raised above the top horizontal wall of the part. This enables column end faces to be through-machined and facilitates nuts tightening. The highest strength and rigidity are possessed by the design pictured in Fig. 73 p in which the vertical wall of the part is matched with the column extreme points. Figure 74 shows flange designs of conical and spherical castings. With small (in respect to flange diameter) cone sizes the flange is connected to the part walls by means of a tulip-shaped mouth Fig. 74. Flanges on conical and spherical parts (Fig. 74a), thus ensuring a streamlined transition of the force flow from the part walls to the flange. In cones of large diameter sizes the joint between the flange and the part walls is reinforced with ribs (Fig. 74 b-d) or by positioning fastening holes in recesses (Fig. 74e, /). With small slope angles the recess walls are extended excessively. In these cases they are given a semi-closed arched form (Fig. 74 g). To obtain the greatest rigidity and strength the walls are position¬ ed at the flange periphery and the recesses ceiling linked with the walls by internal ribs (Fig. 74k). The sizes of the vaulted recesses in height and cross-section must permit easy assembly of fastening parts. With insufficient height (i.e., when the Fig. 75. Mounting fasteners in recesses clearance h between the recess ceiling and stud end face is less than nut height h 0 , Fig. 75a), the unit can be assembled only by raising the part (Fig. 756) and placing all the nuts on the studs ends, a procedure seriously hampering the assembly. The correct version (h > h 0 ) is pictured in Fig. 75c. When using bolts (Fig. 75 d) the recess height must be greater than the bolt length. 5.1. Alignment of Flanges 105 To reduce the weight of low flat flanges the latter are often shaped (in plan), making them narrower in width at the areas between fasten¬ ing bosses (Fig. 76a-c, g-i). In the extreme limit the flange actually vanishes so that only bosses remain which are cast directly onto the part walls (Fig. 76 d, j). These methods must be used very carefully- 4 (a) W (c) Fig. 76. Reducing the weight of cast flanges since the flange rigidity and strength is lowered and the connection between the bosses and part walls weakened. When reducing the flange width to trespass beyond the fastening holes centre line is not recommended (Fig. 76 b, h). It is useful to strengthen the connection between bosses and part walls with local ribs. More preferable are solid flanges (Fig. 76/, l) ensuring higher rigidity and better fasten¬ ing stability. 5.1. Alignment of Flanges Cylindrical flanges are usually aligned by a shoulder and a recess which are machined on the flanges and fit to one another (Fig. 77a, b). In the joints fastened with screwed-in bolts or studs the above-said Fig. 77. Centring of cylindrical flanges recess is substituted by the minimum internal flange diameter of one of the flanges (Fig. 77c). Alignment can also be accomplished on an external flange rim (Fig. lid). The aligning shoulder must be at least 3-4 mm distant from the extreme points of fastening holes (Fig. 78a), otherwise thin rags (m) or sharp-pointed burrs ( n) will form which easily break in service and impair the shape of sealing gaskets. 106 Chapter 5. Flanged Connections In exceptional cases, in order to obtain smaller sizes, the aligning step can be provided at the area where fastening holes are located (Fig. 78 b-d). This method is practicable only for bolted connections; cutting threads in stepped holes and screwing-in fastening elements into such holes offer formidable diffi¬ culties. Stepped holes are machined on assembly which complicates manufactur¬ ing. The designer should not position the centre-aligning step behind the axial Fig. 78. Centring of flanges line of fastening holes (Fig. 786, c) as this would inevitably lead to the formation of sharp rags and burrs. Furthermore, such connections bend as the fastening bolts are being tightened. As obvious from Fig. 78 c a gap formed in the joint accumulates dirt. An acceptable interval at which the centre-aligning step may be positioned is 0.25 hole diameter from the holes axial line (Fig. 78d). To attain close contact between the joint surfaces it is necessary to preclude the corners of aligning surfaces from touching each other. This can be achieved by providing a chamfer that clears the fillet radius where the aligning shoulder meets the flange (Fig. 79a), or it can also be accomplished by radial, face or diagonal undercuts (Fig. 79 b-d). The height H (see Fig. 79a) of centre-aligning steps can be taken for conventional joints at 0.5 Y D {D — centring diameter). 5.2. Machining the End Faces of Fastening Holes 107 A usual mistake made when designing flanged connections is that the aligning recess weakens the flange (Fig. 80a, region 1). R)l Fig. 79. Centring shoulder designs The defect can be cured by making the flange thicker (Fig. 806) or (if sizes and casting technology permit) by reducing the diameter of the centring surface (Fig. 80c). Flanges other than round are made flat, and positioned by fitted pins. Non-critical detachable parts (covers, casings, etc.) are held in place by fastening elements alone. 5.2. Machining the End Faces of Fastening Holes When designing flanged connections the designer must anticipate and specify in the drawing the methods by which the bearing surfa¬ ces under nuts and fastening bolt heads will be machined. Cylindrical flanges may most simply be lathe-turned (Fig. 81a). However, with cast components the lathe-turning adversely affects the strength as it removes the hard surface crust and undercuts the flange in the black surface transition region. It is also inadvisable to lathe-turn flanges with jutting bosses (Fig. 816, c) as the tool receives multiple repeated impacts from the bosses and quickly becomes blunt making it difficult to obtain high standards of accu¬ racy and surface finish. It is more reasonable to machine bosses by milling cutters (Fig. 81d) or core drills, centre-guided by previously made pilot holes (Fig. 81e). The most productive process is machining with a combi¬ nation tool that is a core drill combined with a twist drill (Fig. 81/). In view of the inevitable dimensional inaccuracies encountered in 108 Chapter 5. Flanged Connections moulding practice, the core drill diameter D is made larger than the nominal boss diameter D 0 (on average D = 1.2Z> 0 ). This must be allowed for when specifying joint dimensions. When spot facing sunk bearing surfaces (Fig. 81g) the flange con¬ tour should be 3-4 mm distant from the extreme points of the machin- Fig. 81. Machining methods for nut and bolt head bearing surfaces ed surfaces. Otherwise the possibility of forming the easily broken rags and burrs occurs. If a core drill cannot be used directly on the work surface then inversed core drilling is practised, where the core drill is mounted upon an arbor passed through a previously drilled hole (Fig. 81 h). Machining efficiency is sharply reduced with this method, so that the latter is applicable in the event of holes not less than 10-12 mm in diameter. The end faces of holes in half-closed recesses are machined by milling (Fig. 82a) or inversed core drilling (Fig. 82b). The height and radius R of recesses, taken in cross-section, must be consistent with the dimensions of cutting tools. The modem highly accurate casting techniques (shell mould casting, chill casting, pressure die casting, investment casting, etc.) obviate, in certain cases, machining of support surfaces. However, for critical, heavy-loaded structures 5.2. Machining the End Faces of Fastening Holes 109 machining is obligatory with the assurance that the bearing surfaces are strictly perpendicular to the hole axis, thus avoiding fastening bolt distortion and flexure. Table 6 gives the moulded flange design relationships for the conventional range of bolt diameters (d — 8-20 mm). The machining allowance t depends on the size and grade of accu¬ racy of the casting. On drawings which depict cast components Fig. 82. Machining bearing surfaces in recesses the value of t is generally omitted, which, nevertheless, in no way implies that the designer may neglect this value when calculating the dimensions of parts. The minimum distance s between the machined and the nearest black surfaces is established proceeding from the accuracy standards of casting, dimensions of parts and distances of surfaces from the 110 Chapter 5. Flanged Connections Relationships of Cast Flange Elements Table & Flange elements Mini¬ mum flange height h Mini¬ mum di¬ mensions of machined surfaces (a, R, D/2) Mini¬ mum bolt axis distan¬ ce b from machined wall Minimum bolt axis distance A from flange end face 1.5 d 1.2 d 1 . 2 d 1 . 2 d 1.3d 1.2 d 1.7 d 1.5 d black and machined datum points. With respect to the small- and medium-sized components (200-500 mm) obtained by conventional sand casting s = 3-5 mm; these values can be reduced by 30-50% by applying higher-accuracy casting methods. 5.3. Diameter and Spacing of Bolts The proper choice of the diameter and spacing of bolts depends on many factors, the most important of which are the operational conditions, material of parts and rigidity of structures. The require¬ ments are quite different for connections subjected to small static loads and connections undergoing high cyclic and dynamic loads, working under pressure and where complete sealing is necessary. For the simplest cases (flange-coupled connections, loaded with moderate forces, free of internal pressure and elevated tempera¬ tures), the following approximations can be recommended. 5.4. Three-Flange Joints 111 The diameter of bolts fastening cylindrical flanges d = 6 + (0.015 to 0.018) D where D is the mean flange diameter. The thickness of the flanges: for components made of grey cast iron and light alloys h = 6 + (0.022 to 0.025) D i h l i— h p!-^ j| — u The pitch of bolts (a) Fig. 83. Bolt spacing for flanges of different rigidity for components made of steel or high-strength cast iron h = 4 + (0.022 to 0.025) D j l = ad For small-sized non-rigid flanges (Fig. 83 a) a = 6-8; for mode¬ rately rigid flanges ^(Fig. 83b) a = 8-10; for flanges of higher rigidity fastened with large bolts (Fig. 83c) a = 10-12. The parameters of joints subjected to cyclic loads and operating at elevated temperatures are obtained by calculation (see Chapter 1). 5.4. Three-Flange Joints When designing housing-type components it is often necessary to connect three flanges in one unit. Let us take, as an example, a unit comprising an intermediate partition (membrane) installed between the joint faces of two housings (Fig. 84). The simplest method is tightening the membrane between the flanges of the housings (Fig. 84a-e), the alignment being accomplished from internal or external shoulders. The installation accuracy is the highest in Fig. 84b, c and e (the alignment is effected from one cylindrical surface). The design presented in Fig. 84e is less strong than the others. To 112 Chapter 5. Flanged Connections avoid interference the flange is usually assembled with an axial clearance of 0.1-0.2 mm, owing to which proper membrane tighten¬ ing cannot be assured. Should it be necessary to preserve integrity of mechanisms during dismantling, then it is better to secure the membrane on one of the housings (Fig. 84/-t). After the housings are separated from each (a) (3) (C) (d) ie) (t> (7”) (7J J 1°) (p) Fig. 84. Three-flange connections other the membrane remains secured on one of the housings, carrying all the members mounted on it. Designs differ in alignment methods. In terms of accuracy the most preferable versions are those shown in Fig. 84 h, i. Independent mounting of the membrane can also be implemented by holding (pinching) the flange with various forms of pinch bolts (Fig. 84 j-p), which may be shouldered (Fig. 84;'), with stop rings (Fig. 84fc), tapers (Fig. 84 1) and nuts (Fig. 84m). Occasionally the membrane is secured with the aid of tapped bushes 1 (Fig. 84 n), into which the fastening bolts of the other hous¬ ing are screwed, or the membrane can be held to the housing with independent bolts 2 , 3 (Fig. 84o, p) spaced between bolts 4, 5, which hold the housings together. The heads of type 2 bolts are lodged in the housing flange apertures, while those of type 3 are sunk in the membrane. 5.5. Conical Flange Joints 113 5.5. Conical Flange Joints For butt-jointing pipelines, cylindrical sections and loaded con¬ nections use is made of quick-acting split clamps which tighten upon outer conical flange surfaces (Fig. 85). Parts may be connected at any angle in the joint plane. When angles must be fixed or when the connection transmits torque fitted pins are used. Joined parts are aligned with the aid of cylindrical shoulders (Fig. 85a). Sometimes the joint is made plain (Fig. 85 b) relying upon the alignment given by the clamp’s conical surfaces. The latter method is preferred when the parts to be joined cannot be brought together axially and must be manoeuvred in the joint plane. Conical flange joints ensure a good strong joint with a relatively moderate tightening force being applied to the joint. Assuming the simplest transfer of the tightening force P to two points (as in Fig. 85c, a rigid clamp), then the axial tightening force will be p - 2P ax tan a/2 where a is the cone angle. For the usual a values varying from 20 to 30°, Par = (8 to 10) P. The cone angle a on the clamp collars is made 1-2° less than that of the clamping flanges (Fig. 86a) so that the tightening force is applied closer to the collar base with the aim of increasing joint rigidity and sealing reliability. Similar results are obtained if the clamp walls are made flat (Fig. 86 b). Pictured in Fig. 86c-e are conical flange connections with sealing gaskets. 8—01418 114 Chapter 5. Flanged Connections Figure 86 f-h illustrates various designs of thin-walled tube conic¬ al flange connections. In certain cases inverted conical flange connections are employed. The invert¬ ed conical flanges of the connected parts have through slots (Fig. 87a). During assembly the projecting lug of one flange enters the slot in the other, thereby forming between the flanges a conical cavity into which is placed the central tightening collar (Fig. 876). The binding clamp of a taper flange joint must open fully so that it may be installed on the flanges from the side and axially and Fig. 86. Conical flange connections assure uniform circumferential flange tightening, i.e., be compliant radially. The clamp must be quick-acting. The clamp comprising two halves fastened together by bolts (Fig. 88a) is not quick-acting. A more practicable design has its Fig. 87. Inverted conical flange connection halves linked with a pin and tightened by one bolt (Fig. 88b, c). In the quick-action lock with a swing bolt (Fig. 88 d) radial slits are provided in the hoop walls to enhance the compliance of the system. A circlip-type lock is applied to fasten light joints (Fig. 88e). In a load-carrying structure furnished with a quick-action lock (Fig. 88/) the coupling bolt extends through a cylindrical nut 1 5.5. Conical Flange Joints 115 mounted in a half-open cup; the bolt end terminates in articulated joint 2. As the bolt is being unscrewed, the nut leaves the cup and the bolt can be freely swung away by rotating it about the joint axis. In the design pictured in Fig. 88g the two halves are tightened together by means of a swing lever with pressure screw 3. In the Fig. 88. Collar clamps version presented in Fig. 88k use is made of a swing-type three-link mechanism (“frog”). In the latter case an elastic packing must be inserted between the flanges, which would compensate for the rigi¬ dity inherent in such a mechanism in the locked position. Figure 88i illustrates an elastic clamp made of a steel ribbon with welded-on channel-sectioned segments 6. The coupling bolt extends through pin 4 of the articulated joint and is screwed into the cylin¬ drical nut 5. 8 * Chapter 6 General Principles which Should Be Followed when Designing Units and Parts 6.1. Unification of Design Elements When designing one should repeatedly use elements for all parts of the design which in the assembly process reveal average rated parameters and allow the parts’ list to be reduced to its mini¬ mum. Subject to unification in the first instance are fitted connections (their nominal sizes, type of fits and classes of accuracy), threads (diameter, pitch and classes of accuracy), splined and keyed con¬ nections, fasteners, specifications, etc. It is also advisable to reduce the number of material grades, unify surface finish classes, finishing operations and electroplating techniques, welding methods, forms of welded seams, etc. Figure 89a-c illustrates the arrangement of a typical engineering unit (a shaft carrying some elements and mounted in a bronze bush). In the design presented in Fig. 89a the choice of fitting diameters is badly thought over. The basic fitting size (diameter of the shaft bearing journal), taken from standard specifications ( 50), is cor¬ rectly chosen, but from this point mistakes occur. Wishing to eco¬ nomize on the consumption of scarce bronze, the designer used a bush wall thickness of 3.5 mm, thus obtaining a non-standard size for the external diameter of the bush(<£ 57) and, moreover, while striving to enhance the shaft strength at tbe fitted connection, he reduced the shaft diameter relative to the journal diameter by 3.5 mm on each side, thus obtaining again a non-standard diameter ( 43), which, in turn, meant choosing an M42 screw thread size for the clamping nut. In the arrangement in Fig. 89 b, which is based on standard dimen¬ sions, the external diameter of the bush is 60 mm and the diameter of the fitted connections 40 mm. Hence, the clamping nut thread size is M37. However, standardization of the dimensions in this case has lowered somewhat the shaft strength and also increased the weight of the bronze bush. The most rational version depicted in Fig. 89c has its shaft journal diameter equal to 55 mm, bush O.D. 60 mm and the fitted connection diameter 45 mm. M42 Comparative Table of the Back Gearing Dimensions Table 7 Design element Diameters and fits Design para¬ meter according to Fig. 89d jj| Design para¬ meter according to Fig. 89e Qnty Design element Design para¬ meter according to Fig. 89d Qnty Design para¬ meter according to Fig. 89e Qnty SOP 1 Threads M25 ■ M30 2 35,4/.R l M37 H 40 A/S 1 35P 2 Keys 6 i 40A/P 1 b, mm 8 3 8 40,4//? 2 42,4/S 1 10 1 45 A/S 2 45 A/S 1 Tooth mo- 4 1 5QA/Dh 2 dule, m 5 1 5 1 45A/Dh 1 5QA/R 6 1 1 55A/Dh 1 Total 18 7 6.2. Unification of Parts 119 Naturally, other solutions are possible, but under any circumstan¬ ces it is important for the diameters to be standard. Another example is given in- Fig. 89 d, e (back gear). The design in Fig. 89<2 has significant irregularities in the sizes of fitting diameters, threads, keys and tooth mod¬ ules. In the rational version shown in Fig. 89e the number of fitting dimensions has been reduced and the keys and to¬ oth modules unified. The ne¬ cessary tooth strength for the small gears is obtained by the increased face width. The uni¬ fication results are presented Fig. 90. Unification of wrench sizes in Table 7. All-in-all the num¬ ber of machine elements is cut down from 18 to 7 items. To illustrate the unification of wrench sizes, let us consider a unit for adjusting a reduction valve (Fig. 90). In the design shown in Fig. 90a three wrench sizes (1-3) are used, whereas the unified design in Fig. 90b uses only a single wrench size (4). 6.2. Unification of Parts Individually designed parts must be unified as much as possible. This is particularly important for labour-consuming parts and those in general use (gears, clutches, chain links, etc.). Figure 91a shows a conveyer chain composed of links of two types. The rational design (Fig. 91b) has unified links. The tightening clamp illustrated in Fig. 91c consists of two labour-consuming com¬ ponents. Connection by an intermediate link (Fig. 91 d) enables the clamp to be made of two similar halves. Figure 91e, / shows the uni¬ fication of pressed steel components in the unit of a built-up pulley, while Fig. 91g, h shows unification embodied in the design of a cylindrical pressed reservoir. Often unification is only attained as a result of purposeful, ori¬ ginal, constructional working out of the design. In an angle drive presented in Fig. 92a the required gear ratio is ensured with the aid of two different shafts having different bevel gears. To position the lower gears within preset spaces, one of the gears is displaced relative to the other and the tooth face width of the driving gear 1 made longer. As a result, the design contains five gears (numbered 1 to 5) and two shafts (numbered 6 and 7). 6.3. Principle of Unitization 121 - In the alternative, Fig. 926, the diameter of lower gears 2 is de¬ creased so that they can be driven by the one gear 1. To keep the gear ratio the same the diameter of upper gears 3 is increased. As a result of the conversion the required number of gears is reduced to three (1 to 3) and of the shafts, to one (4). Now, let us examine a reduction gear unit with two concentric shafts running at the same speed in opposite directions (Fig. 93a). The unit driving shaft carries two gears, one of which ( 1) engages with gear 2 of the reduction unit, while the other (3) meshes via idler 4 with gear 5. The total number of gear units is four {1-3, 2, 4, 5). The large number of components and complexity of the design are caused by the necessity to avoid fouling between gears 3 and 5, which means reducing the diameter of gear 3 and, accordingly, that of gear 5 (in order to retain the same gear ratio). In the unique solution (Fig. 936) the design is simplified as fol¬ lows: gear 1 of the drive shaft meshes on one side with the reduction unit right-hand pinion and on the other side, with gear 2 of the drive. The number of gears is thus cut down to two and the pinions and gear wheels of the unit are identical pairs. To realize this design it is sufficient to separate the gear wheels by a distance s sufficient to engage the pinions. 6.3. Principle of Unitization It is advisable to design units as self-contained assemblies, which are individually assembled, adjusted, run-in, tested and finally installed as completely ready-to-operate units upon a machine. Such successive unitization allows parallel and independent assembly of machine units, facilitates mounting, prompts prototype testing and eases the employment on new machines of finished units which were checked during manufacture. Furthermore, unitization simpli¬ fies repair as worn out units can be directly replaced by new ones. Sometimes unitization may complicate the design, but in the long run, it always gives large savings in the total machine manufactur¬ ing costs, reliability and convenience in use. Some examples of the unitization of small assemblies are presented in Fig. 94. In the design illustrated in Fig. 94a the reduction valve is mounted directly in the housing body. When the valve is assembled in a separate bush the design becomes unitized (Fig. 946). The end face seal depicted in Fig. 94c is unsatisfactory because during dismantling the sealing disk 1 is forced by the spring to leave the ways and slots which prevent its rotation, and the unit disintegrates. Equally unsuitable is the seal assembly. The intro¬ duction of locking ring 2 unitizes the construction (Fig. 94 d). Another example of unitization is offered in Fig. 94 e-g where the unit of a distributing slide valve is shown progressively improved^ 122 Chapter 6. General Design Principles The construction given in Fig. 94e is grossly wrong: an accurate hole intended to receive the valve was made directly in the frame -casting. At the spot where the valve is located (i.e., in the concentra¬ tion of material) there may occur pits and voids which make valve sealing impossible. Moreover, as the hole wears out during use it •can only be repaired by introducing bushes. The first step to an improved design is to mount the valve in an intermediate bush (Fig. 94/) made of a high-grade material with a Fig. 94. Examples of unitization good wear resistance. In the most rationally designed version (Fig. 94 g) the component has been fully unitized. Here the unit is made separately; it is assembled, ground-in, verified for geometry and attached to the frame on a milled surface. Shown in Fig. 95a is a worm reducer which is coupled directly with the machine driving shaft. The worm wheel shaft is mounted in supports installed in different housings. To maintain support alignment when machining is difficult. The assembly procedure is 6.3. Principle of Unitization 123 very ambiguous as it is necessary to fit the worm wheel first onto the main shaft, install the reducer housing and then assemble the worm by screwing it into the worm wheel teeth. To verify the cor- Fig. 95. Unitization of gear transmissions rectness of engagement and align the worm wheel under these cir¬ cumstances is very difficult. In the unitized construction (Fig. 95b) the worm wheel shaft is mounted in two supports, one of which is accommodated inside the housing and the other in the housing cover. Both supports can be machined together, thus obtaining good alignment. The shaft end 124 Chapter 6. General Design Principles is coupled with the driving shaft by the splined adapter 1. Hence, the reducer assembly is considerably simplified. The gearing mounted in a frame (Fig. 95c), in addition to the draw¬ backs inherent in the non-unitized constructions, presents diffi¬ culty in introducing the idle gear shaft into the support housing. As soon as the housing is removed from frame, it loses both supports and falls to pieces. To verify the correctness of meshing and align¬ ment of the shaft and supports is impossible. In unitized constructions the gear supports are mounted in diaph¬ ragm 2 (Fig. 95 d), or in bracket 3 (Fig. 95e) screwed by its feet to the gearing casing. With the latter design it is easy to inspect the mechanism during assembly. Figure 95/, g shows the unitization of a gear drive mounted on a frame; Fig. 95 h, i, a reduction unit driven from an engine crankshaft. In the unitized version (Fig. 95i) the reduction unit is mounted in a separate housing; the shaft torque is transmitted by means of adap¬ ter 4. 6.4. E limination of Adjustments Fitting and adjustment of units and parts when in position must be eliminated. Adjustments, particularly when accompanied by machining or fitting operations, lower assembly productivity and deprives the construction of interchangeability. An example of fitting in situ is shown in Fig. 96a, b. The gear is placed in position on the shaft so that it meshes with its mating gear. This done, the position is fixed by a self-piercing screw (Fig. 96a) or by a taper pin (Fig. 966), which requires drilling and reaming in situ. Chips inevitably fall into the unit. After the machining it is necessary to dismantle the unit, flush it and then reassemble. Marking during assembly with a subsequent transfer to machine tools further complicates assembly. A better engineering method is to lock the gears in position by ring-shaped stop washers which are placed into grooves previously machined on the shaft (Fig. 96c). If a bearing has been installed in a housing by the “in situ” tech¬ nique (Fig. 96 d), its originally correct position will be disturbed every time the bearing is dismantled, thus necessitating readjust¬ ment. Fixing the bearing by dowel pins (Fig. 96e) entails machining on assembly. The correct solution is displayed in Fig. 96/, where the bearing is aligned against the bore in the housing, this bore being machined to the required accuracy, which ensures correct functioning of the unit. If a straight machine way is mounted on a bed as illustrated in Fig. 96g, then its position must be verified on assembly and holes drilled to suit fastening bolts. In this case the way will not be gua¬ ranteed against offset within the clearance values existing between 126 Chapter 6. General Design Principles the bolts and their holes. Fixing by dowel pins (Fig. 96 h) means drill¬ ing and reaming holes in the way and bed. In the rational design (Fig. 96i) the way is mounted in groove machined in the bed. The gear drive illustrated in Fig. 97a is unsatisfactory. The gear supports are bolted to the housing. The fitter is compelled to adjust the support position so as to assure correct gear meshing. Dismantl¬ ing will inevitably disturb the arrangement, thereby necessitating readjustment on assembly. The supports can be fixed in position by dowel pins (Fig. 976), but this will involve additional job opera¬ tions in assembly. In the correct design (Fig. 97c) the gear studs have been aligned against holes whose spacing relative to each other has been maintai¬ ned within close tolerances during machining of the housing. In the best version (Fig. 97 d) all the gears are enclosed inside a common housing, which brings about full unitization and advantageous ope¬ rating conditions. Figure 97e gives a wrong, and Fig. 97/, a correctly designed ver¬ sion of a back gearing with a V-belt drive. 6.5. Rationalization of Power Schemes The perfection of a structure, its weight, size and, to a significant degree, its capability of work are dependent on the rationality of its power scheme. A power scheme is rational if all acting forces are mutually equalized over as short a section as possible with the help of elements loaded mostly in tension, compression or torsion (but not flexure). As an example let us examine a screw-type conveyer (Fig. 98 a) driven from an electric motor via a worm-type reduction unit 1 and chain transmission 2. The conveyer housing, which is several metres long, is made of steel sheet and rests upon four tubular legs. The main error consists in the fact that the housing is loaded with the drive force (the direction of force is shown by an arrow), which bends and deforms the non-rigid housing, mounted on unsteady sup¬ ports. Owing to a small internal clearance between the conveyer screw and housing the screw edges scrape the inner walls. The inten¬ sified friction increases the driving torque which, in turn, increases the bending force still more and, consequently, the friction and finally the screw jams in the housing. The deffect can partially be eliminated by changing screw rota¬ tion with a corresponding change in screw helix. The driving side of the chain drive is now the lower side of the chain drive and the torque moment acting upon the housing is substantially reduced. Alternatively, it is possible to move the reducer into the symmetrical plane of the entire assembly, redistribute the legs, strengthen the housing and mount it upon a stiff foundation. Nevertheless, all 6.5. Rationalization of Power Schemes 127 these measures fail to eliminate the principal defect, namely, the presence of external forces in the system. A radical solution is presented in Fig. 98ft. Here the screw is driven from a flange-mounted electric motor via a reduction unit installed coaxially on the housing end face. The drive torque and reactance? Fig. 98. Improvement of power scheme of screw-type conveyer torque on the housing quench each other at the connection joint. The housing is absolutely free from the action of external forces and deformation is eliminated. In the drive of a suspended conveyer (Fig. 99a) comprising a reduc¬ tion unit 1, bevel gearing 2 and spur gears 3, which impart motion to the driving sprocket 4 of a chain drive, the load-carrying elements are irrationally arranged. Indeed, the support units of the drive, the fastening bolts and foundations are all loaded with drive forces and a large number of elements are in flexure. The drive units are 128 Chapter 6. General Design Principles separated; they are installed on individual bases and are not fixed one relative to the other. Such a design only ensures satisfactory functioning after the mechanisms have been carefully aligned. Furthermore, the cast iron gears are not protected from fouling ■and can only be periodically lubricated by packed grease. The Pig. 99. Improvement of power scheme of suspension-type conveyer drive supports of the horizontal and vertical shafts are also lubricated periodically. The entire unit is excessively large, explained by the isolation of units and also by the employment of a low-strength material (cast iron) for manufacture of the most critical components (gears). In fact, the design is typical of old design practice. In the modern unitized enclosed construction with a rational power arrangement (Fig. 99 b) the drive is from a flange-mounted electric motor, installed vertically, which obviates the necessity for the angular transmission. The required gear ratio is obtained in one co¬ axially mounted reduction unit. The employment of gears made of a high-grade and appropriately heat-treated steel in conjunction with the application of a centralized oiling system, sharply reduces overall sizes. The drive forces are absorbed in one casing. The casing and base are loaded only by the force of the final driving sprocket. These changes result in tremendous savings in the weight and size of the installation, simplicity of manufacture, convenience in mounting and attendance, reliability and a long service life. 6.6. Compensators 129 6.6. Compensators For multistation machine systems having mechanical drives the design of couplings which transmit torque is of extreme importance. Such a coupling must compensate for lengthwise displacements, (a) (b) (c) Fig. 100. Misalignments in coaxial connections parallel misalignment e and conical misalignment a of the units being connected (Fig. 100). For this purpose involute splined connections are most commonly employed (Fig. 101). The advantages of involute teeth, are as fol¬ lows: due to the thickening of the tooth towards the root (particularly with positive correction) the tooth posses¬ ses increased strength and stress concentration at the root is not great; involute teeth, both external and internal (with a sufficiently large rim dia¬ meter), can be machined to close tolerances on conven¬ tional gear cutting machi¬ nes. External involute teeth Fig. 101. Compensators with involute teeth can be given high surface hardness by heat and chemical-heat treatments, followed by fi¬ nish machining on conventional gear-grinders. The operating conditions for teeth in compensating connections are much more harder than for the centring splined connections. The enhance their compensation ability such connections are made with an increased circular clearance s = (0.05 to 0.07) m, where m is the tooth module. In the case of a misalignment the forces are con- 9-01418 130 Chapter 6. General Design Principles centrated at the tooth extreme edges lying in a plane perpendicular to the misalignment. As a result the linear contact over the tooth face changes to point contact, thus sharply increasing local bear¬ ing stresses. Since each tooth passes through the zone of loading twice per revolution, the load upon the teeth will be cyclic, regardless of torque behaviour. The working capacity of the connection can materially be inten¬ sified by enhancing surface hardness of teeth. To eliminate work har¬ dening and promote removal of heat, produced during tooth impacts and warpage, ample lubrication is obligatory. The most effective Fig. 102. Determination of maximum conical misalignment in a compensator method of obtaining an efficient connection is to increase the gear rim diameter since this offers certain advantages in manufacture, namely, the internal teeth can be gear-cut in place of using expensive broaches. The amount of misalignment admissible in a coupling is limited, first of all, by the tooth edge contact in a plane perpendicular to the misalignment direction (Fig. 102a). The teeth positioned in the misalignment plane have much displacement' freedom as the clear¬ ance in the radial direction with a standard pressure angle (20°) is approximately three times greater than the circular clearance. The maximum possible misalignment angle a may be determined from the relation tan a = ~f where s = circular clearance in teeth l — tooth face width The maximum displacement of the compensator’s extreme points S = L tana^- (6-1) where L is the compensator length. 6.6. Compensators 131 Compensation can be improved by decreasing the tooth face width, obtained without weakening the tooth by increasing the gear rim diameter. The circumferential force acting upon the splined rim n_ D ( 6 . 2 ) where Mtg e = torque transmitted D = mean diameter of the splined rim For small displacements the tooth strength is determined from the bearing stress on the spline faces Gbear where l P Izh P Izam (6.3) = spline width 2 = number of splines h = am = working spline height, proportional to the spline tooth module a = constant Since z = Dim, then P 2M t „ Dla Gbear whence l = 2 M tge D2la const D 2 Obea D2 (6.4) The spline length l [Eq. (6.4)1 and maximum possible displa¬ cement S of the compensator’s extreme points [Eq. (6.1)] as a function of diameter D are shown graphically in Fig. 103 (s and l when D = D t are as¬ sumed to be equal to unity). Fig. 103. Spline length l and maximum displacement S vs. spline diameter To decrease the load upon the tooth edges and increase the permis¬ sible misalignment angle, it is good practice to give a barrel-like shape to the teeth. The tooth end face edges should be rounded along the entire tooth contour. For large conical misalignments the tooth faces and roots should be made spherical (see Fig. 1026). Illustrated in Fig. 104 is an intermediate compensating bush which permits large free displacements and in its general scheme resembles a cardan joint. The bush has incomplete internal tooth rims; the toothed portions of both rims are at right angles. The disks with external teeth have complete gear rims, which ensures correct assembly at any angular attitude of the flange relative to the inter¬ mediate part. 9 * 132 Chapter 6. General Design Principles Some of the design alternatives of compensating connections are presented in Fig. 105. Connection by means of splines cut directly on drive shafts (Fig. 105a) is unreasonable, because the compensa¬ tion ability of such a joint is rather low and it is determi¬ ned only by the amount of spline displacement in the gap between spline edges. In¬ creasing the drive shaft shank length (Fig. 1056) only worsens the position since the spline d shank owing to inevitable manufacturing inaccuracies and assembly is apt to runout proportional to the distance at which the end is spaced from the drive shaft supports. If a splined reducing bush is freely mounted upon the splines of shafts and acts as an adapter (Fig. 105c), then the compensating ability, determined by the total clearance in the splines will be twice that of the design given in Fig. 105a. Fig. 104. Cardan-type compensator The best construction has the driving ends of the splined bush spaced apart longitudinally (Fig. 105 d). Compensation in this instan¬ ce is improved due to the displacement of the bush itself. Still better is the case when a longer bush is used (Fig. 105e). However, this alternative is more difficult to produce because of the different splined rims on the bush ends. The best design has a long splined shaft, a torsion bar (Fig. 105/), having high compensation characteristics. 6.7. Torsion Bars 133 6.7. Torsion Bars Torsion bars (torsion bar springs) not only compensate for mis¬ alignments, they also damp torque-induced vibrations, thus contri¬ buting to quieter and smoother operation. This advantage is par¬ ticularly important in machines operating under pulsating torsional loads (piston-operated machines). Thanks to their small radial dimensions, torsion bars can readily Fig. 106. Torque transmission by means of torsion bars be included in shaft interiors, thereby adding to the compactness of design (Fig. 106a). To avoid the breakage of torsion bars by overloading, provision is occasionally made of a torque limiter. The latter is generally made in the form of a splined bush mounted concentrically with the tor¬ sion bar (Fig. 106b). The side clearance in the torque limiter splines exceeds that of the torsion bar splines so that the limiter takes up the load as soon as a certain angle of torsion is reached. General-purpose torsion bars are made of silicon spring steels, grades 60C2A, 70C3A or 60C2XA, which, when subjected to appropriate heat treatment (harden¬ ing plus moderate tempering), possess a fatigue limit t 0 = 65-70 kgf/mm* under pulsating torsional loads and x_ x = 30-35 kgf/mm 2 under symmetrical alternating torsional loads. 134 Chapter 6. General Design Principles For stressed structures, as well as for structures operating under elevated temperatures, use is made of silicon-nickel, silicon-tungsten and silicon-vana¬ dium steel grades 60C2H2A, 65C2BA and 60C2XOA, for which t 0 = 80- 90 kgf/mm 2 and t-j = 40-50 kgf/mm 2 . The elastic torsion of torsion bars can be enhanced by increasing the design stress level. For pulsating cycles t = 30-40 kgf/mm 2 is generally adopted, which corresponds (in terms of fatigue limits) to a safety margin of the order of 1.5-2. In structures designed for a limited service life the admissible stresses can be increased to 80-100 kgf/mm 2 . The fatigue strength of torsion bars can materially be improved by means of plastic strain hardening. For instance, torsion bars intended for operation under alternate cyclic loading, are hardened by shot blasting, while torsion bars, designed for pulsating loads are hardened through prestraining (i.e., applying a static moment of the same direction as the operating one, the stress level being 20-40% in excess of the material yield point). The shot blast technique and prestraining procedures enable torsion bar service life to be approximately doubled. The best results are obtained by means of strain hardening (shot blast¬ ing with prestrain) which further increases service life by 20-30%. Torsion bar splines are strengthened by rolling in the axial direction with rolls profiled to suit the spline tooth space. Involute splines are strengthened by rolling against a hardened sizing gear. The torsional stress in a torsion bar i = kgf/mm 2 (6.5) vv tor8 where M tqe = transmitted torque 1 Ytors — 0.2 d 3 (1—a 4 ) = torsion resisting moment of the tor¬ sion bar section a = d 0 /d = ratio between the I.D. and O.D. of the bar section (for solid torsion bars 6 a = 0) The torsion bar twist angle is

17 ^ Eboit Fnut-Ftottj^-j— If the nut and bolt are made of same material (E nUt = E bo n) IFnut^Fboitj^ ( 6 . 11 ) Substituting in Eq. (6.11) the values F bo it= 0.785d 2 and Fnut — 0.785 X X(D 2 — (P), where d is the bolt diameter and D running external diameter of the nut, we get _ “-tVth i6 - ,2 > When x = h, diameter D = oo. Obviously, this condition can only be approximated. For any finite nut diameter (where thejthread plane finishes) the upper threads will be loaded more than thej lower ones. Uniform distribution of thread loads can be achieved by increasing the bolt end compliance. Let a conical recess be made in the bolt [stem (Fig. 1346) so that its peak begins in the same plane as the thread. The running diameter 0 of the recess is h where d. is the initial recess diameter (assumed to be equal to that of the bolt). 6.17. Effect of Elasticity upon Load Distribution 169 The running cross-sectional area of the bolt is ^bolt = 0.785(d2-02) = 0.785d2 [V— (-f-) 2 ] Substituting this relation in Eq. (6.11) gives = (6.13) When x = h, diameter D = d\/~3 = 1.73 d. This nut shape can be realized in practice. The cross-section of the nut above the threaded portion is determined from the equal strength condition for the nut and bolt in tension Fnut = Fbolt (6.14) Let the diameter of the recess in the nut above the threaded portion be 6, Then, according to Eq. (6.14) 0.785 [(1.73d)2—02J = O.785d2 whence 0 = dty r 2 = 1.41d Figure 134c shows a design approximation of the theoretical nut and bolt shapes, where the load is uniformly distributed along’the threads. Figure 135 illustrates how a heavily loaded screwed connection (namely, a propeller blade fastening unit) can be rationally redesign¬ ed. The factor of elasticity must necessarily be taken into conside¬ ration when designing bearing units. Figure 136a, b shows a paired installation of antifriction bearings. In Fig. 136a the greater part of the load is taken up by the bearing positioned in the ri¬ gidity node (housing wall plane). The other bearing mounted at the hub end is lightly loaded because of the hub being compli¬ ant. The bearing loads can be evenly distributed (thus contri¬ buting to a better load-carrying capacity of the unit as a whole) if the hub end is stiffened with another web (Fig. 136b). Another example of elasticity utilization is given in Fig. 136c in which the bearings are installed in a thin-walled steel bush. Thanks to the elasticity of the bush the system can adapt itself to shaft distortions, and resembles a system which has bearings mounted in a spherical support. Thus, by utilizing elasticity one can obtain optimum distribution of load between bearings. In the bearing unit loaded with radial blade fastening 170 Chapter 6, General Design Principles force P and unilateral axial load Q (Fig. 136d) it is better to distri¬ bute the load according to the bearings function: one of the bearings should be loaded with the radial force and the other, with the axial Fig. 136. Effect of housing elasticity on load distribution in paired bearing installation force. This can be realized by mounting the bearings in an elastic cantilever-type bush. Bearing 1 located in the rigidity node (i.e., in the fastening flange plane), takes the radial load. Bearing 2, mounted at the cantilever end receives axial load only. 6.18. Fitting to Several Surfaces It is good practice to avoid, whenever possible, fitting parts to several surfaces (Fig. 137a, b). As a rule, parts should be fitted to one 6.19. Tightening up on Two Surfaces 171 tion, elastic deformations, thermal expansion of the system or under compression of sealing gaskets. Rough errors, like those sketched in Fig. 137a, c, are only charac¬ teristic of inexperienced designers. Still more typical errors include superfluous fitting or alignment, etc. For example, fitting a prisma¬ tic loose key throughout the entire contour of the keyway (Fig. 137/) will heavily, and, which is still more important, quite unnecessarily complicate the production. It is more reasonable to fit the key to the working edges only, providing clearances on the key end faces, as well as between the key top surface and the keyway bottom (Fig. 137g, h). 6.19. Tightening Up on Two Surfaces '"Occasionally it may be necessary, due to some design features, to^tighten a unit of two surfaces. Figure 138 (tightening three flan¬ ges) reviews techniques applied under these conditions. Simultaneous and uniform tightening of all surfaces (Fig. 138a) presumes] that the end flange faces should be machined together Fig. 138. Tightening against two surfaces. Fastening of intermediate flange and fitted or have undergone extremely accurate manufacture. If a flange protrudes from its recess (Fig. 1386), an interference may occur and the tightened part will bend. On the other hand, if the flange is'sunk in its recess (Fig. 138c), the flange axial fixing will be lost. The introduction of elastic gaskets (Fig. 138d-/) improves the design and seals the joint if the gasket is sufficiently thick and flexible and overlaps non-flush sealing surfaces. Whenever the good sealing and accurate axial fixing of a flange are necessary, it is recommended to employ gaskets of some soft metal (red copper, lead, aluminium, etc.) whose thickness exceeds the depth of the recesses where the gaskets are to be fitted. During 172 Chapter 6. General Design Principles tightening the metal gasket plastically deforms (Fig. 138g), seals the joint and fixes the flange. Some space must be provided to enable excessive metal to freely flow out. The bearing stresses arising in the gasket under the action of working axi¬ al forces must be below the yield point of the gasket material. Otherwise, accurate axial fixing will be lost. If harder metals (brass, bronze, low-car¬ bon annealed steel) are used for the gask¬ ets, these must be made corrugated or comb-like (Fig. 138ft) to assure plastic defor¬ mation. Spring gaskets (Fig. 138i, /) are also applied. Figure 139 shows two-piece crank conne¬ cted in a main journal by means of trian¬ gular-shaped end-face splines, an antifri¬ ction bearing being simultaneously clamped between the two parts of the crank. jThe tightening of the splines and inner bearing race in this case is ensured on the account of the deformation of metallic rings 1 placed on either side of the inner race. 6.20. Axial Fixing of Parts Parts must be fixed axially at One point so that they can freely self-align throughout their length. If, for instance, a pin is fixed by self-tapping screws at both supports (Fig. 140a) then additional stresses are likely to occur as a result of thermal dimensional changes. In the correct design shown in Fig. 1406 only one end of the pin is fixed, while the other end is free to move in its support. In the badly designed herringbone gearing (Fig. 140c) the lower gears are fixed in position twice: by their teeth and stops m. To obtain the complete coincidence of fixing positions is impossible. The error can be eliminated by providing clearances s, which assures gear self-alignment by their teeth (Fig. 140<2). The gear shaft mounted in plain bearings (Fig. 140e) is fixed in two points spaced far away from each other. Accurate fixing in this case is impossible because a large clearance must be provided between fixing surfaces in order to avoid seizure of bearing surfaces due to thermal expansion of the housing. Further¬ more, due account must be taken of the inevitable inaccuracies of manufacture and assembly due to the required clearances between mounting surfaces. The design is somewhat improved by bringing the fixing surfaces closer together (Fig. 140/). In the correct designs (Fig. 140g, ft) the shaft is fixed over a short distance (in the design presented in Fig. Fig. 139. Tightening against several surfaces. Two-part crank journal 174 Chapter 6. General Design Principles 140 h it is fixed practically in one plane); the other end of the shaft is allowed to self-align in its support. Free areas of parts must be given certain margins for self-align¬ ment and compensation for manufacturing dimensional inaccura¬ cies. Let us consider the case of a shaft installed inside a housing on antifriction bearings. In the design illustrated in Fig. 141a the axial dimensions, predetermining the mutual location of the shaft, bear¬ ings and housing, are held to the nominal size. In the alternative Fig. 141. Introducing margins on fitting surfaces pictured in Fig. 1416 the following clearances are provided: m — on the fitting surface of the housing to suit the floating bearing; h — on the fitting surface of the housing relative to the fixing stop rings; k —in the thread for the fastening nut; n —on the fitting surface of the shaft to suit the floating bearing. The clearance values are found by calculating the dimensional chains and thermal expansions in the system. The maximum clear¬ ances should be provided in the areas of transitions to black cast surfaces where dimensional variations are particularly great (for medium-sized castings of moderate accuracy such clearances are taken at 3-4 mm). 6.21. Control of Direction Parts reciprocating linearly along two guideways should be guided along one way only, the other guide only supporting the part (Fig. 1426, d). A simultaneous guiding along two ways (Fig. 142a,c) means much more stringent machining requirements to the guides and ways. Any change in temperature conditions may impair orientation and, hence, the part may seize in its guideways. 6.22. Mounting Surfaces 175 If the use of two guideways is inevitable, their manufacture must be simplified as much as possible. In a design with two guiding rods (Fig. 143a) the necessity to accurately keep the centre distance bet¬ ween the rod sockets in the driven part and the guiding holes in the mmmmm (a) J ^\\\\\\\W WMMM/M ( 6 ) (c) Fig. 142. Guiding of parts (o) and (e) wrong; (6) and ( d) correct (d) housing can be eliminated by increasing the diameter of the sockets (Fig. 1436) and fixing the rods in their sockets after aligning them with the guiding holes. Another method consists in machining the Fig. 143. Ensuring accurate guidance over two surfaces rod sockets and guiding holes together. In this case the diameters of the guiding holes and the rod sockets must be the same (Fig. 143c). However, different fits for the rods are required, namely a sliding fit in the housing and an interference fit in the rod sockets. 6.22. Mounting Surfaces Generally speaking, any mounting surface for detachable parts should be flat. Fastening on cylindrical surfaces should be avoided (Fig. 144a). The manufacture of such connections is labour consuming. The fitting surface of the detachable component must be machined in a fixture to assure the equality of the diameters of the mounting surfaces on the part and housing and it is rather difficult to uni¬ formly tighten angularly-spaced bolts. The correct design with a flat mounting surface is pictured in Fig. 1446. Sometimes it may be necessary to fasten parts to surfaces position¬ ed at an angle to one another (Fig. 144c) .J The connection seems strong and rigid. Its manufacture, however, is difficult as it is 176 Chapter 6. General Design Principles absolutely necessary to keep the angular equality between the fitted part and housing to preclude the deformation of the part during fastening. Fastening bolts must be alternately tightened and each time by small amounts in order to assure the close fitting of the part to both mounting surfaces. Obviously, a design with a plane mount¬ ing surface is much more preferable (Fig. 144 d). Flat attachment is of particular importance when air-tight con¬ nections have to be obtained. Sealing surfaces must be free of any Fig. 144. Clamping on curved surfaces steps, as well as of outer or inner corners. Fitting to curvilinear sur¬ faces is inadmissible. An example is presented in Fig. 144e, which shows a cover enclosing an angular space in a housing. The design has two defects. First, it is practically impossible to seal off the space end walls forming reentrant angle a. Secondly, the proper tightening of the cover is made impossible by the fact that screwing up one row of bolts will interfere with screwing up the other row of bolts. The latter of these two defects is eliminated in the design pictured in Fig. 144/ where the cover is tightened up by a row of diagonally arranged bolts. This technique is rather common in fastening shields over spaces which do not require sealing, and also when enclosing through tunnels. Should the air-tightness be necessary, the only correct solution will be that illustrated in Fig. 144 g where the cover is fitted on a flat surface. Also erroneous is the design of a cover intended to enclose a hatch in the corner of a welded sheet-steel housing (Fig. 144A), since it is practically impossible to provide close tightening over a curvi¬ linear surface even when using a thick gasket. A correct design sketched in Fig. 144i incorporates a flat frame welded to the hatch and forming a flat fitting surface. 6.23. Butt-Jointing on Intersecting Surfaces 177 6.23. Butt-Jointing on Intersecting Surfaces Butt-jointing on intersecting surfaces complicates the manu¬ facture and sealing of joints. An incorrect joint design is pictured in Fig. 145a. Here the side cover 1 is mounted upon the joint of the housing and the upper Fig. 145. Fitting on intersecting surfaces cover, which means that its fitting surface must be machined after the assembly of the housing and the upper cover, and to assure the tightness of the joint, a thick elastic gasket must be used. In Fig. 146. Methods of fastening split housings the correct design shown in Fig. 1456 the housing-to-upper-cover joint is moved beyond the side cover. A non-technological design is that of a housing made up of two halves joined in vertical plane AA (Fig. 145c). The upper cover is mounted on the joint of the halves. Still worse is the design in Fig. 145c? where the cover is joined with the housing halves in two mutually perpendicular planes. In the correct design (Fig. 145e) the joint surfaces are separated. Figure 146a shows a rotor machine housing split in the horizon¬ tal plane with two end covers also split horizontally, which makes the joint sealing difficult. The slightly improved version in Fig. 1465 has solid end covers. The best designs have either a housing and covers (Fig. 146c) joined in vertical planes (axial assembly), or a housing (Fig. 14 6d) split in the horizontal plane (radial assembly). 12-01418 178 Chapter 6. General Design Principles 6.24. Interchangeability of Rapidly Wearing Parts It is good practice to manufacture any rubbing or rapidly wearing part of a machine component as an easily interchangeable piece. The latter can be made of a material with special properties which are absent in the basic material of the component. Figure 147a illustrates a T-slot made in the body of a cast-iron bed. To provide for machining convenience, longer service life Fig. 147. Introduction of replaceable parts and changeability it is advantageous to make the T-slot as a sepa¬ rate piece from some durable material and to fasten it to the bed, e.g., in a slot (Fig. 147b). To ensure reliable operation of the threaded pair in an inter¬ nally threaded steel rod for a lead screw (Fig. 147c), it is advisable to use a bronze bush (Fig. 147d) which possesses antifriction pro¬ perties and can easily be changed. Figure 147c illustrates an erroneously designed seal with split spring rings fitted in shaft grooves, the external cylindrical sur¬ faces of the rings sealing against the housing. When the grooves in the shaft and the hole in the housing get worn, both these expen¬ sive components have to be rejected. In the correct design (Fig. 147/) the rings are assembled in a changeable bush and work in a sleeve made from a harder material. When installing an antifriction bearing in a light-alloy housing (Fig. 147g) the seating surface rapidly crushes during operation. Machining the hole even slightly oversize causes the rejection of 6.25. Accuracy of the Alignment of Parts 179 the entire housing unit. It the correct design (Fig. 147 h) the bear¬ ing is installed in an intermediate bush made from a harder mate¬ rial, which reduces the seat wear and enables defective housings to be rectified. To install the valve of an internal combustion engine in the manner depicted in Fig. 147i, i.e., directly in a cast iron head, is bad practice. It is better to install the valve in a guide bush made of a harder material (Fig. 147;) and introduce an insert-type seat of a heat-resistant material. Figure 147 k-m shows a water-cooled engine block. If the cylinder working surfaces are machined directly in the cast block (Fig. 147/c) the design is bad. To obtain high-quality mirror-finish bores in a large casting is difficult. Furthermore, the cylinder walls may have defects (blisters, voids, pits, etc.) which sometimes/are exposed only during finish machining. Clearly, a single defective cylinder means rejection of the whole block. Finally, excessive wear in one of the cylinders during operation means failure of the entire expensive unit. The correct solution is to use insert liners (Fig. 1471). The best design has wet liners which are directly surrounded by water (Fig. 147m). This system has significant additional advantages: simp¬ ler casting techniques, reduced weight and better cooling of the cylinders. 6.25. Accuracy of the Alignment of Parts Parts requiring accurate mutual location are preferably installed in one housing with a minimum number of transitions and fits. Fig. 148. Reduction valve design For example, let us take a reduction valve unit (Fig. 148). The most important factor predetermining the reliability of the unit is the fitting of the valve tapered chamfer on the seat. In our exam- 12 * 180 Chapter 6. General Design Principles pie it is accomplished with a number of transitional connections, each being the source of inaccuracies, namely: running fit between valve stem 1 and guide bush 2; press fit between bush 2 and cover 3\ slide fit between cover 3 and casing 4\ press fit between valve seat 5 and casing 4. The design requires accurate alignment of the following elements: in the valve—guide surface and bevelled edge of the plate; in the bush—bore and seating surface; in the cover—hole and centring shoulder; in the casing—centring bore for the cover and hole for the seat; in the seat—chamfer and seating surface. Fig. 149. Angle gear drive When being ground-in to its seat, the valve is centred in guide bush 2. The seal obtained by the grinding-in procedure is spoilt during reassembly because of the displacement of cover 3 in relation to casing 4. In the rational design (Fig. 148 b) the valve is centred directly on the seat. The accuracy of the valve motion is now dependent on a single joint only, i.e., on the slide fit of guide shank 6 in seat 7 of the valve. In this case strict alignment of the following elements will assure good functioning of the unit: in the valve—guide surface of the shank and chamfer; 6.26. Relief of Precision Mechanisms 181 in the seat—chamfer and seating surface. The remaining elements can safely be manufactured to lower stan¬ dards of accuracy. In the grinding-in process the valve is centred in its seat and reassembly does not destroy the seal obtained by the grinding-in process. Another example is an angle gear drive. In the design shown in Fig. 149a the gears are installed in different housings 1 and 2, which hampers the manufacture. The joint surface of hous¬ ing 2 must be machined strictly parallel to the pinion axis. Never¬ theless, the mounting accuracy is disturbed when the sealing gasket in the joint is tightened. Another defect is the obstructed access to the gears. The axial position of the gears can be adjusted only by blacking-in in several repetitive tests, the gear being detached each time. The adjustment accuracy may be impaired during reas¬ semblies due to a non-uniform tightening of the gasket. With the gears installed in one housing, their positional accuracy is not disturbed during installation and subsequent reassemblies (Fig. 1496). The gears can now be easily inspected during assembly* The gears’ engagement adjustment is greatly simplified. 6.26. Relief of Precision Mechanisms Precise moving connections and mechanisms should be relieved from the action of external forces which may cause excessive wear or disturb the correct working of the mechanism. The working surfaces when in use must be protected against extraneous forces and negligent treatment. Figure 150a illustrates a conical plug-type cock with a handle provided directly on the plug shank, which is wrong for the follow¬ ing reasons: the effort exerted in turning the handle is taken up by the cock ground-in surfaces; occasional impacts against the handle may spoil the sealing surfaces. Furthermore, an inexperienced operator may pull the cock handle axially, thus injuring the tight¬ ness of the cock. The self-alignment of the plug in the conical seat is hampered by the fact that the shank should simultaneously be centred in the cock cover. In the design presented in Fig. 1505 the driving shaft with the handle are mounted in a separate body and spline-connected with the plug. Here the plug is relieved from external forces and can self-align in its seat. This design has an additional improvement: the axial position of the plug is adjusted by pressure screw 1. At the beginning the screw is loosened, allowing the plug (actuated by a spring) to tightly fit its seat. Then the screw is slightly screwed in, thus slightly lifting the plug. This hardly impairs the tightness, but enables the cock to be turned much easier. As the plug wears down, the adjustment is repeated. 182 Chapter 6. General Design Principles In the design of a flat face slide valve illustrated in Fig. 150c the tightness may easily be spoilt by depressing inadvertently the handle and thus displacing the slide valve away from the surface being sealed. The defect is eliminated if the driving shaft and the valve are separated (Fig. 150d). In the unit of a distributing cam-operated slide valve (Fig. 150e) the actuating forces from the cam are taken up by accurately ground-in surfaces of the slide valve. During operation these susfaces Fig. 150. Relieving mechanisms from extraneous forces wear down. In the better alternative (Fig. 150/) the drive forces are taken up by an individual tappet, hence the slide valve is relieved from the action of the transverse forces and the wear on the sealing surfaces is reduced to the minimum. Another example is the drive of an internal combustion engine valve. In the version presented in Fig. 150g' the cam acts directly on the plate which is screwed into the hollow valve stem. As the valve opens, the cam runs against the plate and the valve distorts (within the guide clearance); the sealing chamfer of the head leaves its seat and forms a crescent-shaped slit. This is particularly dan¬ gerous with exhaust valves because a jet of hot gases gushes through 6.27. Coupling Parts Made from Hard and Soft Materials 183 the slit, causing a one-sided erosion and burning of the valve. When the cam runs off the plate, the valve rests sideways in its seat. Consequently, the wear of the valve sealing chamfer and seat occurs on one side. In the design shown in Fig. 150/i the transverse force components are taken up by intermediate sleeve 2 , so that the valve is acted upon by an axial, centrally-applied force. The increase in the mass of the reciprocating parts means some limitation to the engine speed. This defect is eliminated in the design depicted in Fig. 150t where the valve is driven through intermediate lever 3. Now the valve is practically relieved from the action of transverse forces (not as fully as in Fig. 150 h). 6.27. Coupling Parts Made from Hard and Soft Materials Frictional units incorporating parts made from hard and soft materials must be designed so that the rubbing surface of the harder and more wear-resistant material overlaps completely the rubbing surface of its mating member made of a softer and less wear-resistant material. Observation of this rule assures a uniform wear of the Fig. 151. Combination of parts made from materials of different hardness softer part. If the soft surface overlaps the harder one, a stepped worn spot appears which impairs the operation of the unit. Let us now examine this fact through examples. In the end trun¬ nion bearing installed in a bronze bush (Fig. 151a) the trunnion end is shorter than the bush. After wear, part l of the bush will be step¬ ped and hinder the axial self-alignment of the trunnion. It is also incorrect to hold the axial dimensions to the nominal sizes, for manufacturing errors, assembly inaccuracies, as well as thermal deformations in the system may produce a displacement of the trunnion end inward of the bearing with the same result as previ- 184 Chapter 6. General Design Principles ously. In the correctly designed version (Fig. 1516) the trunnion projects beyond the bush for a distance, which allows all possible variations of the longitudinal dimensions to be catered for. In the self-aligning spherical footstep bearing (Fig. 151c) the rubbing surface diameter of the steel disk is less than that of the bronze support. Eventually the disk will inevitably wear steps in the bronze support, thus obstructing the shaft self-alignment. The correct design is shown in Fig. 151d. A wrongly designed face-type mechanical seal is pictured in Fig. 151e, in which an immovable textolite! bush 1 is pressed by springs 2 against a hardened steel disk 3 rotating together with Fig. 152. Design versions of conical stop cocks the shaft. Here the frictional surface of the steel disk is smaller than that of the textolite bush, thus causing uneven wear. In the correct design (Fig. 151/) the steel disk face overlaps the textolite bush. Figure 151 g, h shows respectively wrong and correct designs of a cylinder-piston unit. In the tappet unit in which the tappet slides in a bush (Fig. 151i) the oil-distributing recess is provided in the tappet stem. The more reasonable design has the recess machined in the bush (Fig. 151/), thus ensuring equal wear of the stem and bush. Figure 152a-c illustrates a stop-cock whose conical plug, made of a hard material, is mounted in the body of a soft material. The design in which the plug is shorter than the conical socket in the body (Fig. 152a) is wrong as during the grinding-in process a step is formed at the part h of the socket. This step hinders the deepening of the plug into the socket. The same occurs as the socket wears in operation. In the improved version (Fig. 1526) the plug end extends out of the socket, thus ensuring uniform wear. An insignificant step, however, may appear at part m of the plug. The best design is given in Fig. 152c, where the plug top is slightly below the socket surface. Now the wear of the socket and plug in no way hampers the deepening of the plug. The design in operation has a self-grinding-in property. For the inversed case, i.e., a soft plug and hard body, the afore¬ said remains valid. The construction in Fig. 152d is incorrect. 6.27. Coupling Parts Made from. Hard and Soft Materials 185 - because from the grinding-in processes and wear a step is formed at part n. This step prevents the plug from going deeper. The defect can be eliminated by lowering the plug below the socket face (Fig. 152e). A still better design is shown in Fig. 152/, where any possi¬ bility of step formation is utterly excluded, both on the plug and in the socket. Thus, we may generalize all cases, including those when the plug^ and body are of equally hard material, and say that the plug top end must be below the surface and the lower smaller end must protrude beyond the socket. This statement remains true for fixed connections made of mate¬ rials which differ in hardness. It would be wrong to fit a soft huh Fig. 153. Tightening of tapered joints onto a conical portion of a steel shaft (Fig. 153a). Here the hub front face, i.e., the one nearest to the nut, extends beyond the shaft taper. As the hub is ground-in to the taper, and also while the hole loses its true roundness in running, a step is formed at part h. This step obstructs the hub assembly on the shaft in the course of repeated tightenings. In a correct version (Fig. 153b) the shaft taper extends beyond the hub, so that changes in the hole size when fitting and because of bearing deformations in no way interfere with the proper tightening. Should the material of the hub be harder than that of the shaft (a fact rare in practice), then the most dangerous case is the one when the hub rear face does not reach the start of the taper (Fig. 153c), because now, in the process of grinding-in and tightening, a step is formed at part m of the shaft. In the correct version (Fig. 153d). the hub end overlaps the taper. Conical fits fail to assure accurate longitudinal fixation. The mutual posi¬ tion of the fitted parts depends to a great extent upon the manufacturing accu¬ racy of the tapers both on the shaft and in the hub, upon the tightening force, number of reassemblies, bearing deformations and wear of the surfaces. For 186 Chapter 6. General Design Principles these reasons conical fits must not be used when accurate axial positions are required. For example, let us consider the pinion carrier of a planetary drive, whose disk is attached to the casing by means of the satellite shafts. In the design presented in Fig. 153e it is practically impossible to maintain precisely distance l in respect to all fixture points. Because of the inevitable inaccuracies in the taper diameters and axial distances between them, the lengthwise disk displacements in the course of tightening will differ from shaft to shaft. This displaces, buckles and overstresses the disk. Furthermore, it is difficult to maintain strict centre distances between the tapers. It is also impossible to ensure registration of holes in the connected parts by machining them together (as is often done with cylindrical holes). In fact, the parts cannot be accurately as¬ sembled. The design showing shafts with only one tapered end (Fig. 153/) is better only because the errors are halved. Such kinds of assemblies should have cylindrical fitting surfaces with the disk tightened up on stops (Fig. 153g). The distances between the fixing steps on the shafts can be held to close tolerances. Registration of the holes’ centres in the cover and casing is attained either by machining the holes in a jig or by through machining them together. 6.28. Elimination of Local Weak Spots Local weak spots due to decreased cross-sectional areas and stress ■concentrators can sharply reduce the strength of components. Often such a weakening results from a miscalculation of the cross- sectional areas of the part. This mistake often occurs when designing small parts deemed not worth calculations, which allegedly take on loads (except tightening forces). A characteristic example is illustrated in Fig. 154a. A nipple has a hole 0 20A. To enable free retraction of cutting tools the hole bottom has an undercut 020.5. The threaded nipple portion (M24-1.5) also has an undercut 022 to clear the thread-cutting tool. For the given dimensions the wall thickness at part h (over the internal groove) is 1.25 mm, and at part m (under the external groove) the wall thickness is 1 mm. Clear¬ ly, the nipple wall will break even in the case of a weak tightening against the hexagon. In the improved design (Fig. 1546) the diameter of the part across the internal groove section is increased to 26; hence, the minimum 6.28. Elimination of Local Weak Spots 187 wall thickness is increased to 2.75 mm. A larger thread M27 with a finer pitch (s — 1 mm) is used, which increases the wall thickness at the external groove to 2.75 mm. Very often a local weak spot can be eliminated by shifting the weakening element into a zone of larger cross-sections. The example given in Fig. 155a illustrates this. The bush is sharply weakened by two grooves in the same plane (position h), which Fig. 155. Elimination of local weak spots enable a free withdrawal of cutting tools. Moving the internal groove into the flange plane (Fig. 1556) strengthens the part. In the nozzle cap nut (Fig. 155c) the undercut should be provid¬ ed in the hexagon plane (Fig. 155<2). The internal nut (Fig. 155e) is seriously weakened at the section m where a tool retraction groove is provided and at the sections n where spanner-receiving holes are drilled. In the improved design (Fig. 155/) the holes are replaced by splines and at the weak section the wall thickness is enhanced. If a lever is secured to a shaft in the way shown in Fig. 155g, the hub will be seriously weakened by the keyway. In the better design (Fig. 1551i) the keyway is moved to the area of increased cross-section, namely, to the transition where the hub joins a longi¬ tudinal rib. An eye bolt loaded with a tension force P (Fig. 155i) is weakened by hole d located in the most severely stressed area and intended to receive an oiler. In the strengthened construction (Fig. 155/) the hole is arranged in the thickened portion of the eye. 188 Chapter 6. General Design Principles The bolt head with an internal hexagon produced by broaching (Fig. 155fe) is heavily weakened by a groove at the section /, which is intended to clear the broach where the bolt stem turns into the head (this section undergoes bending). Changing the broaching to upsetting (Fig. 155Z) eliminates the weakness and the bolt acqui¬ res additional strength thanks to a more favourable grain orien¬ tation. The strength of the leaf spring fastening unit (Fig. 155m) is lowered by the screw-fastening hole at the critical section. The strengthening methods in Fig. 155w-p complicate the manufacturing procedure and cause greater waste since the springs are produced by blanking from steel sheet or ribbon. The strongest and most technological design of the spring made from steel strip and secured by strap v is shown in Fig. 155g. Now the spring takes up stresses with its unweakened section and is safely protected by the strap at the most critical section. 6.29. Strengthening of Deformable Areas The strength of a part can be substantially enhanced by reinforc¬ ing its non-rigid areas which are readily deformed under the action of operating forces. Let us take, for example, the slot-tang joint unit of shafts illu¬ strated in Fig. 156a, b. In the irrational design (Fig. 156a) the tang of the driving shaft deforms the forks of the slotted shaft when trans¬ mitting torque. This loosens the fit. A press-fitted binding hoop over the fork end (Fig. 1565) sharply increases the joint strength and rigidity. In the universal joint (Fig. 156c) the torque is transmitted by a pin press-fitted into the driving shaft knuckle and passing into slots machined into the driven shaft end. In the reinforced construc¬ tion internal slots are used instead of the forked end in the solid shaft (Fig. 156d). The fork-end connection in Fig. 156e is irrational, because the fork opens under the action of tension forces (shown by light arrows). The unit strength will considerably improve, if the forks are addition¬ ally tightened on to an auxiliary bush (Fig. 156/). Figure 156g shows the fastening section of turbine blades mounted in a T-shaped tenon on the rotor rim. Under the action of centrifugal forces the blades roots move apart (shown by dashes). In the better design (Fig. 156/i) the roots are provided with spurs m entering the annular recesses in the rotor rim, thus preventing the roots from mov¬ ing apart. The two-part crank pictured in Fig. 156i is coupled in such a way that its crankpin tightens against the flat web face n. The elastic 6.30. Composite Constructions 189 deformations of the unit from the working loads cause conical defor¬ mation of the crankpin end, and work hardening and welding of the fitting surfaces. Fig. 156. Reinforcement of units In the strengthened design (Fig. 156/) the crankpin end is coned and fits into the conical recess l in the web. The tightness of the fit¬ ted collar stops the conical deformation and work hardening. 6.30. Composite Constructions In many cases it is advisable to separate parts, connecting them correctly later by press-fitting, welding, brazing, riveting, etc., or by means of fastening bolts. Such composite designs simplify machin¬ ing and geometry, save weight and economize on rare or expensive materials. The separating into difJerent components in many instances saves considerable labour when manufacturing large bed castings. A bed casting with the lower halves of longitudinal shaft bearing housings cast integrally (Fig. 157a) is not correct technologically, as it is necessary to bore the cylindrical holes in the bearing caps and bed simultaneously and maintain strict parallelism of the hole centres, etc. The machining is particularly difficult when the bearings arranged in a line are spaced at substantial distances. In the design with separate housings (Fig. 1576) machining is much easier, as it only means milling or planing the surfaces to which 190 Chapter 6. General Design Principles the bearing housings are attached. The housings are positioned on the bed by dowel pins. Another example is the mounting of rollers on the corners of a mo¬ vable bed (Fig. 158a). To obtain axial alignment of the rollers and Fig. 157. Bed with bearings position them in the same horizontal plane is very difficult, especial¬ ly when the rollers are spaced rather far from each other. The use of separate eye fastenings simplifies the processing (Fig. 1585). Fig. 158. Simplification of machining of housing-type components through introduction of detachable parts The best version is given in Fig. 158c. Here the seating surfaces for the roller lugs can be through-milled, which ensures that the rollers are positioned in the same plane. Furthermore, the lugs are more rigid than the eye fastenings and are reliably fixed against rotation relative to the bed. Easing the machining of slots in beds through the application of detachable parts is illustrated by the examples in Fig. 158 d-g. Figure 159 illustrates how fabrication by welding can simplify machining. The cluster gear (Fig. 159a) is very difficult to produce because of the shaped cavity between the gears. The composite design (Fig. 1596) in which two separate gears are joined by resistance Fig. 160. Composite constructions as substitutes for forgings 192 Chapter 6. General Design Principles welding enables easy machining of the gears of required configuration. The angle lever which is rather difficult to blank (Fig. 159c) can advantageously be changed to a welded design composed of two simi¬ lar parts of simple shape (Fig. 159d). Other examples of composite constructions are given in Fig. 159e-;\ Composite constructions are extremely helpful for intricately shaped components. The cross-piece (Fig. 160a) which serves to transmit motion from a cam to two valve tappets can be produced only by stamping in enclosed dies, the latter being economical only in the event of batch production. In piece production it is reasonable to use a built-up component made from two lathe-turned parts press- fitted together (Fig. 160b). Composite constructions are popular substitutes for forgings when dealing with shafts having large-diameter flanges, etc. Figure 160c-e shows an example of how a forged flanged shaft is transformed into a welded construction. In Fig. 160 d the problem is solved but partially: the shaft here is produced from round rolled stock and involves difficult drilling and boring operations. In the improved design presented in Fig. 160e the shaft is produced of a seamless pipe; the shaft end is formed by welding a length of a smaller-diameter seamless pipe. Such a construction reduces ma¬ chining operations to the minimum. It is very essential to note that the employment of composite constructions as substitutes for forgings is justifiable only under small-batch production conditions. Generally it is more advantageous to produce parts from forgings that approximate as close as possible in configuration, as this results in substantial strength gains, lesser machining, greater productivity and, in the final analysis, less cost. However, even in batch production it is sometimes more reasonab¬ le to resort to composite designs as a means of simplifying forgings. Let us take, for instance, the shaft of a planetary reduction unit with a pinion carrier (Fig. 160/). Clearly, an attempt to forge these units as one will cause quite a lot of difficulties. In the composite design (Fig. 160g) the pinion carrier is made as a separate part and fitted to the shaft on splines. The forging of both parts is very much simplified. In a number of cases components are made of light alloys, which enables the weight of the unit to be radically decreased. The cam plate (Fig. 161a) is better made from aluminium alloy, to which a steel cammed rim is attached (Fig. 161b). Other examples are: the axial compressor rotor with a steel drive splined rim (Fig. 161c, d); the brake drum with a steel liner on its frictional surface (Fig. 161e, /). Composite constructions find wide application as a means of eco¬ nomizing on expensive and scarce materials. Fig. 161. Composite constructions made of light alloys Fig. 1G2. Composite constructions as a means of saving scarce materials 13-01418 194 Chapter 6. General Design Principles Worm gear teeth (Fig. 162a) working in sliding friction conditions are normally made of antifriction bronze which is rather expensive. To economize on bronze, bronze hoops are used (Fig. 162a), which are then press-fitted upon carriers made of some cheaper material (e.g., carbon steel, cast iron). Compared to Fig. 162a the amount of bronze in Fig. 162fc is reduced to one third. Clearly, with the gears shown in Fig. 162c only the gear rim need be made of high-grade steel (Fig. 162d). Figure 162e illustrates a rod made completely of alloy steel, with its spherical end being hardened. The more economical construction in Fig. 162/ has only the spherical end made of alloy steel. Lead bronzes and babbits surely belong to short-supply materials. Whenever conditions permit, low-lead or even leadless bronzes and babbits, as well as adequately qualitative substitutes, e.g., antifric¬ tion aluminium alloys should be used. Should the application of difficult-to-obtain non-ferrous alloys be avoided, their consumption must be reduced to the minimum. Let us take, for instance, a housing with many friction surfaces (central bore and lug holes). In the alternative presented in Fig. 162g the housing is made completely of expensive antifriction bronze, but the rationalized version, illustrated in Fig. 162 h uses cast iron (or some other abundant material), and bronze bushes for the friction surfaces. An example of how a material in short supply can be spared in the production of bearing bushes is given in Fig. 162t and /. Here the thick bronze bush (Fig. 162i) is changed to a thin-walled one rolled from strap brass (grade JIC-59-1) and strengthened by rolling and expanding in the seating bore (Fig. 162/). An effective way of economizing on lead babbits is by decreasing the thick¬ ness of babbitting. These days the thickness is reduced to 0.2-0.3 mm (instead of 1-3 mm practised quite recently). The most economic bearings are those with a double-layer babbitting and consisting, in effect, of a babbit layer whose thickness varies within several hundredths of a millimetre. The layer is deposited by an electrodeposition technique upon a sublayer of porous bronze. The deposi¬ tion of babbit in the pores of the bronze sublayer assures good cohesion between babbit and bronze and creates in the bronze sublayer a structural interface which is rather similar to lead bronze in terms of antifriction properties. 6.31. Shoulders Shoulders (Fig. 163a-d) serve as positive rests for components in fixed joints or as stops which restrict the axial displacement of parts in movable connections. The most rationally designed shoulders are those that have (owing to their shape) equal resistance to bending (Fig. 163d). Such shoulders possess the least weight and are simple to produce. The non-working surfaces of shoulders are sloped at 45° 6.31. Shoulders 195 so that they can easily be turned by means of a bull-nose lathe tool whose nose angle is 45°. Shaped shoulders are uneconomical because they are difficult to produce (Fig. 163ft). Shoulders must be kept as low in height as possible: the higher the shoulder, the greater the metal wastes in the form of chips and Fig. 163. Shoulder designs the more difficult the production process. Figure 164 shows how a high shoulder in a gear unit can be eliminated. Shoulder m is for holding the gear shaft in the housing and fixing the gear axially (Fig. 164a). Itis the latterfunc- tion which needs such a high shoulder. Now, replacing the shoulder by the washer n which works upon a low step of the shaft (Fig. 164ft) makes the shaft manufacture easier. Illustrated in Fig. 165a-o are various techniques enabling one either to make lower sho¬ ulders or to eliminate them at all (for fixed joints only). In Fig. 165 b-d the fitted part is tightened against an intermediate washer resting against a step or shoulder of reduced height. Shoulders are often replaced with ring stops having rectangular cross-sections (Fig. 165e). The strength of the unit can further be enhanced by fitting the ring into a cylindrical recess made in the part or in some intermediate member, thus preventing the ring from loosening and leaving the recess (Fig. 165/, g). A very strong stop is ensured by rings enclosed in a conical recess either in the part or in an intermediate washer (Fig. 165ft-/). Fig. 164. Reduction of shoulder height 13 * 196 Chapter 6. General Design Principles Occasionally use is made of half-rings placed into grooves on the shaft and locked either by means of a recess in the part to be tighten¬ ed up (Fig. 165fc, l) or by means of an embracing ring (Fig. 165m). In the design alternative given in Fig. 165rc embracing ring 1 is Fig. 165. Reducing the shoulder height and replacing shoulders fixed on half-rings by locking spring ring 2. The joint is easily disassembled manually by merely shifting the ring axially. In the version pictured in Fig. 165o the shoulder is formed by swag¬ ing a plastic metal ring into a groove on the shaft. The process is done on rotary swaging machines; after swaging, the ring is machined together with the shaft. It should be noted that grooves weaken shafts, and fitting rings into such grooves is not recommended for joints subjected to high cyclic loads. In some cases weakening can be avoided by making the shaft thicker across the section where a groove is needed (Fig. 165/). 6.32. Chamfers and Fillets Generally all external corners must be chamfered (Table 8) and all internal corners, filleted (Table 9). Conventionally the chamfers are made at 45°. The leg c of a cham¬ fer for usual cylindrical components is obtained from the relation¬ ship c — 0.1 D, where D is the diameter of the cylinder. The va¬ lues obtained with the aid of this relationship are rounded off to standard values: c = 0.2; 0.5; 0.8; f; 1.2; 1.5; 1.8; 2; 2.5; 3; 3.5; 4; 5 6.32. Chamfers and Fillets 197 Free, non-contacting surfaces are given chamfers within 0.1-0.2 mm. Such chamfers are not indicated in drawings (in contrast to the design chamfers); a note merely says “Remove sharp comers”. Often the necessity for removing Fig. 166. Overlapping of fillets sharp corners is indicated in the Technical Specifications pertinent to the part, the chamfer sizes and tolerances being specified. Figure 166 shows the methods of overlapping fillets. The overlap¬ ping is effected by a fillet whose radius is larger than that of the part being encompassed (Fig. 166a), or by a recess (Fig. 166fc) or else by a chamfer (Fig. 166c). The latter method is the simplest one. Chamfers Table 8 Design Design wrong correct wrong correct Table 8 ( continued) Application Purpose To facilitate screwing of nuts and threaded rods Threaded connections To facilitate use of spanners To form an annular bearing surface (with diameter £>). To prevent point contact on bearing surface Fillets Table 9 Design wrong correct Application Purpose Design wrong correct Application Purpose Cb vT R1 m B ■ p ■I m Heat- treated parts Parts sub¬ jected to chemical and thermal treatments and thermal diffusion saturation To prevent edges from overheating and decarboni¬ zation. To re¬ duce quenching stresses at tran¬ sitions To ensure uniform satu¬ ration of sur¬ face layer with introduced ele¬ ments Ml I mm,■ Electro¬ plated parts To prevent local variations in current den¬ sity. To ensure uniform depo¬ sition of metal Painted, varnished and poly¬ merized parts To ensure uni¬ form coating. To prolong ser¬ vice life of coatings Cast parts To ensure uniform solidi¬ fication of me¬ tal in cooling. To reduce shrin¬ kage stresses Application Purpose Stamped parts To improve metal flow and fill die reentry angles Parts stamped from sheets To improve metal flow. To prevent disrup¬ tions at transi¬ tion areas Reser¬ voirs To eliminate corrosion at corners. To fa¬ cilitate flushing Parts subject to heat erosion To prevent edges from overheating and burning Table 9 ( continued) Application Purpose Heat-dis¬ sipating fins To improve heat transfer from part body to fins Decorati¬ ve parts To improve outer appearan¬ ce. To facilita¬ te polishing Hand controls To protect hands from in¬ jury and facili¬ tate manipula¬ tion. To facili¬ tate polishing Design Design Table 9 (continued) Application Purpose Any loaded part To increase static and cyc¬ lic strength at transitions Parts subject- To reduce ed [to contact edge pressures loads Machined parts To enhance durability of cutting edges Index Accuracy of parts alignment, 179 Alignment of flanges, 105 Axial fixing of parts, 172 Bending, elimination and reduction of, 137 Butt-jointing on intersecting surfaces, 177 Cambering, 161 Chamfers, 196 Combining design functions, 150 Compactness of design, 146 Compensators, 129 Composite constructions, 189 Connection(s), centring, 79 conical flange, 113 design rules for, 80 glued, 78 loaded, 12 pressed, design of, 67 pressed, with electro-deposited coatings, 65 press-fitted, 37 screwed, centring in, 92 serrated, 77 strength of, 38 three-flange, 111 tightened, 7 unloaded, 7 Control of direction, 174 Coupling of parts of different hardness, 183 Coupling(s), floating cross-sliding, 136 Design, compactness of, 146 of equally strong parts, 152 general principles of, 116 of pressed connections, 67 of screwed connections, 94 Elasticity, effect on load distribution, 164 Elimination, of adjustments, 124 of deformations due to tightening, 143 of local weak spots, 186 Equistrength, 152 , Fillets, 196 Fitting to several surfaces, 170 Fit(s), conical, 75 drive, 37 press, 37 selection of, 46 Interchangeability of rapidly wearing parts, 178 Methods of controlling preloads, 32 Mounting surfaces, 175 Press-fitting, with cooling of parts, 64 with heating of parts, 64 Rationalization of power schemes, 126 Relaxation, 25 Relief of precision mechanisms, 181 Self-alignment, 157 Shoulders, 194 Strengthening of deformable areas, 188 Tightening up on two surfaces, 171 Torsion bars, 133 Unification, of design elements, 116 of parts, 119 Uniform loading of supports, 156 Unitization, principles of, 121 TO THE READER Mir Publishers welcome your comments on the content, trans¬ lation, and design of the book. We would also be pleased to receive any suggestions you care to make about our future publications. Our address is: USSR, 129820, Moscow, 1-110, GSP, Pervy Rizhsky Pereulok, 2, Mir Publishers Printed in the Union of Soviet Socialist Republics OTHER BOOKS FOR YOUR LIBRARY MECHANISMS IN MODERN ENGINEERING DESIGN. BY I. ARTOBOLEVSKY A handbook in six volumes for engineers, designers, and inventors on the mechanisms used in modern engineering design. Each mechan¬ ism is represented by a diagram of its working principle, with a concise description, and classified by a scheme proposed and deve¬ loped by the author. Vol. I. Lever Mechanisms. Illustrates and describes 912 lever mechanism. Contents. Elements. Simple lever mechanisms. Jointed lever mechanisms. Vol. II. Lever Mechanisms (in two parts). Illustrates and describes 1376 lever mechanisms. Contents. Link-gear mechanisms. Slider-crank mechanisms. Lever-cam mechanisms. Gear-lever mechanisms. Lever-ratchet mecha¬ nisms. Flexible-link lever mechanisms. Elastic-link lever mechan¬ isms. Wedge-lever mechanisms. Lever-screw mechanisms. Vol. III. Gear Mechanisms. Illustrates and describes 689 toothed, friction and cam mechanisms. Contents. Simple gear mechanisms. Lever-gear mechanisms. Pin- gear mechanisms. Cam-gear mechanisms. Ratchet-gear mechanisms. Worm-gear mechanisms. Complex gear mechanisms. MACHINE TOOLS. BY N. CHERNOV A textbook for technical colleges and apprentice engineers on the standard parts and mechanisms, electrical and hydraulic equipment of machine tools, the equipment of automatic transfer lines, and N/C machine tools. Contents. Introduction to machine tools. Standard parts and mechan¬ isms. Electrical equipment. Hydraulic equipment. Automatic (Numerical) control. Standard lathes. Turret lathes. Backing-off lathes. Multihead machines. Semiautomatic and automatic lathes. Drilling machines Boring machines. Milling machines. Dividing heads. Principles of tracer milling machines. Thread-cutting machi- ines. Shapers and planers. Broaching machines. Grinders. Lapping machines. Gear-cutting machines. Machine tools built from standard units. Automatic transfer lines. Electrosparking, ultrasonic and other machines. N/C machine tools. Modernization. Installation. Testing, maintenance and repairs of machine tools. MACHINE ELEMENTS. BY V. DOBROVOLSKY et at. A textbook for university mechanical engineering faculties. Contains rules, methods and standards for designing the elements on the basis of their working conditions. Acquaints the reader with the design of individual elements, and the various forms of assembly and drive used in modern engineering. Contents. Part I. Fundamentals of designing machine elements. Criteria of operating capacity and calculation of machine elements. Selection of material. Standardization. Production soundness. Part II. Joints. Types of joint. Riveted joints. Welded joints. Joints formed by interference fits. Threaded joints. Cottered fastenings. Keyed, splined and keyless joints. Part III. Power transmission. Types of drive. Friction drives. Belting. Gearing. Helical and hypoid gears. Worm gears. Globoidal gears. Reduction gears. Chain drives. Power screws. Part IV. Shafts , hearings and couplings. Shafts and axles. SJiding- contact bearings. Rolling-contact bearings. Couplings and clutches. Part V. Springs and frames. Springs. Machine frames. ENGINEERING TECHNIQUES FOR ANALYZING STRENGTH AND RIGIDITY. BY G. GLUSHKOV A monograph on the theory of moments developed by the author, and its application to engineering design problems on the strength, rigidity, and stability of beams and columns. The approximate (and occasionally exact) solutions obtained can be employed to analyze loads of all kinds. The method is also applicable to various-non-linear problems. Contents. Fundamentals of the theory of moments. Application of the theory. Fundamentals of the moment-operational techniques. FORMULAS FOR DESIGNING COMPLEX FRAMES. BY G. GLUSHKOV I. EGOROV, V. ERMOLOV. A collection of formulas for calculating the moments and reactions at supports and joints of complex frames and arches under various loads. Contents. Frames with stepped suppors (16 variants)l Multispan and multistorey frames (49 variants). Arches (14 variants)..Force diagrams and load factors (117 variants). Appendices. Examples of analyses. Formulas for calculating shearing forces, bending moments, and the angles of rotation and sag of simple beams (117 variants).