31
This article was downloaded by:[2007 King Mongkut's University of Technology Thonburi] On: 31 October 2007 Access Details: [subscription number 780012218] Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK Drying Technology An International Journal Publication details, including instructions for authors and subscription information: http://www.informaworld.com/smpp/title~content=t713597247 HEAT AND MASS TRANSFER IN IMPINGEMENT DRYING Suna Polat a a Procter & Gamble Company Winton Hill Technical Center 6250 Center Hill Avenue, Cincinnati, Ohio Online Publication Date: 01 January 1993 To cite this Article: Polat, Suna (1993) 'HEAT AND MASS TRANSFER IN IMPINGEMENT DRYING', Drying Technology, 11:6, 1147 - 1176 To link to this article: DOI: 10.1080/07373939308916894 URL: http://dx.doi.org/10.1080/07373939308916894 PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf This article maybe used for research, teaching and private study purposes. Any substantial or systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Heat and Mass Transfer in Impingement Drying

Embed Size (px)

Citation preview

Page 1: Heat and Mass Transfer in Impingement Drying

This article was downloaded by:[2007 King Mongkut's University of Technology Thonburi]On: 31 October 2007Access Details: [subscription number 780012218]Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Drying TechnologyAn International JournalPublication details, including instructions for authors and subscription information:http://www.informaworld.com/smpp/title~content=t713597247

HEAT AND MASS TRANSFER IN IMPINGEMENTDRYINGSuna Polat aa Procter & Gamble Company Winton Hill Technical Center 6250 Center HillAvenue, Cincinnati, Ohio

Online Publication Date: 01 January 1993To cite this Article: Polat, Suna (1993) 'HEAT AND MASS TRANSFER INIMPINGEMENT DRYING', Drying Technology, 11:6, 1147 - 1176To link to this article: DOI: 10.1080/07373939308916894URL: http://dx.doi.org/10.1080/07373939308916894

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf

This article maybe used for research, teaching and private study purposes. Any substantial or systematic reproduction,re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expresslyforbidden.

The publisher does not give any warranty express or implied or make any representation that the contents will becomplete or accurate or up to date. The accuracy of any instructions, formulae and drug doses should beindependently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings,demand or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with orarising out of the use of this material.

Page 2: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

DRYING TECHNOLOGY, 11(6), 1147-1176 (1993)

HEAT AND MASS TRANSFER IN IMPINGEMENT DRYING

Suna Polat Procter & Gamble Company Winton Hill Technical Center 6250 Center Hill Avenue Cincinnati. Ohio 45224

-and Drying of Paper and Textiles; Impinging lets

In indusoial drying applications, efficient transfer of heat and mass between a drying medium and the material being dried is very critical for the overall economics of the operation. Impinging jets are therefore widely used for their enhanced transport characteristics, especially for drying of continuous sheets of materials such as paper and textiles. In such applications, a thin sheet of material, as wide as 6m in cross machine direction, speeds at velocities as high as 90 kmhr under high velocity jets emerging from a confining surface parallel to the material surface. Many variables and effects need to k considered for proper design of such impinging jet systems: nozzle geometry and sizc. nozzle confieuntion. location of exhaust wns . nozzle-to-surface se~aration. jet-to-je~ separation. &ss flow. jet c x ~ t veloc~ty abd surface motion. For pcrmcnhle mntenals, add~tional cnh~nccmenl of hcat and mass transfer thst occur when some of the mptnglng gas 1s rcmovcd through the mdrenal makes this option an attmctivc one.

Here, we review the above effects and offer predictive correlations from literature which may be used in the design of high velocity impinging jet systems.

In drying of solids, imponant mechanisms that may affect drying are: - Heat transfer . Capillary uanspon of moisture to the surface . Diffusion in liquid phase

Diffusion in gas phase

copytight 0 1993 by Marcel Dckkcr. Ins.

Page 3: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

Phase equilibrium between the gaseous,

liquid and solid phases

Depending on the initial and the final moistures of the material being dried, as

drying proceeds, the controlling mechanism can change from pure heat transfer to some

combination of the above mechanisms. For example, drying of unbound water is a

heat mnsfer controlled phenomena. During this regime, generally called the "constant

drying rate period", drying rate increases with the heat wnsfer rate. Drying of bound

water, on the other hand, is generally limited by phase equilibrium and transpon

phenomena insidc the material. During this latcr stage, the drying rate continuously

decreases as the name "falling ratc period" implics. At thc constant drying rate period,

the temperature of the material stays nearly constant - at thc adiabatic saturation

temperature of the drying medium - and sharply increases during the falling rate period.

If the product to be dried is temperature sensitive, drying conditions at this stage must

be adjusted carefully to avoid product degradation.

If the product is in a form that is amenable to direct exposure to hot gases, heat and

mass transfer rates at the material surface can be enhanced significantly using

impinging jets. Impinging jets are thus particularly useful for drying unbound moisture.

However, should the drying continue to rrmove a significant portion of the bound

moisture, additional advantage of impinging jets is the potential for fine control of local

transfer rates, not only by varying flow ratc and AT, but by a number of geometric jet

parameters: nozzle type, jet diameter (or width), jet-to-surface distance, jet-to-jet

separation and configuration. This many design parameters complicates the design, and

requires more careful fabrication of the equipment to avoid undesired non-uniformities.

At the same time, it provides flexibility: with prior knowledge of these effects, a dryer

can be designed in several stages where drying rate and product temperature can be

controlled to achieve the best product quality.

In industry, impinging jets are used to dry materials in the form of continuous

sheets, i.e. paper and textiles, photographic film, veneer and carpets, or in the form of

granules or pallets, i.e. food and pharmaceulical products. In some cases hot jets are

insened in a bed of particles moving on a conveyor belt, fluidizing the particles.

Tnnspon propenies at particle surfaces in such systems differ from those produced by

impingement and thus will not be pan of this paper.

Page 4: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING 1149

With enhanced transport characteristics of impinging jets, large drying duties can be

achieved in smaller size equipment. Because large volumes of air or hot gases are used

for rcasons of thermal efficiency, gases have to be recovered and reci~ulated. This

adds new considerations to the design, i.c. confinement and location of exhaust pons.

Cost increases because of more complex fabrication and increased air handling systems.

Dryer designs are based on heat and mass transfer measurements done in laboratory

equipment. Although most industrial uses are with systems of multiple jets. most

laboratory investigations have been with single jets. In a multiple jet system when the

jets are spaced widely and therefore not interacdng, use of single jet data is expected to

be valid. However, when the jets are spaced closely and therefore highly interacting.

special multiple jet data must be used in the design. Knowledge of geomeuic and flow

conditions at which jets in a multiple impinging jet system no longer behave as

independent single jets is required to make the distinction between the non-interacting

and interacting jet systems. n e case of multiple jets thus introduces the design

variables like jet-to-jet separation, exhaust flow location and cmssflow.

In drying of continuous sheets of materials, an added variable is the machine speed

relative to velocity of jets. For example, in impingement drying of paper (e.g. Yankee

dryer), the paper moves at speeds as high as 90 km/hr under impinging jets with nozzle

exit velocity of about I00 mls. With large changes in boundary layer conditions at a

moving surface, it is expccted that transfer rates would be significantly different than

those measured in laboratory studies with a stationary surface.

Again, for drying of permeable continuous sheets such as paper or textiles.

impingement drying rates can be increased funher by drawing some of the hot gases

through the product. This type of drying then combines two very effective modes of

heat and mass m s f e r mechanisms: impingement and through flow. Because only a

fraction of hot gases is drawn thmugh the pmduct, this combination drying can be

applied to even low permeability materials as opposed to pure through flow drying

which is only practical to use for drying of highly permeable webs.

This many design variables both simplifies and aggravates the design of an

impinging system. On one hand, understanding and quantification of these multiple

effects - individually or in combination - is critical for an accurate design. On the other

hand, it seems that a wrong choice for cenain design variables can be compensated by

Page 5: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

adjusting others. The difficulty is to find the optimal system design that will satisfy the

requirements of both the product and process simultaneously.

In this presentation, the effects of various design variables on impingement heat

transfer are discussed, and design correlations from literature are given. The equivalent

mass transfer coefficients can be obtained using analogy beween heat and mass

transfer. The intent is to provide a practical framework for the design of impinging jet

dryers.

Heat transfer coefficient, h, is a measure of the heat transfer efficiency. It is

commonly expressed as heat transferred per unit time per unit area per degree

temperature driving force (AT). The AT on which "h" is based naturally affects its

value. Hence, it is very imponant that the AT used between the reporting studies and

the applications should bc consistent. For impingement systems, AT is the difference

between the impingement surface temperature and a reference fluid temperature. fl,- T,). The reference temperature. T,, can be the nozzle exit temperature, Ti, a film

temperature based on the jet and impingement surface temperatures, e.g. (T,+T.) 01

~T,+zT,), or the adiabatic surfacc temperature. 3

Impingement heat transfer is actually a three temperature problem, i.e. in addition to

the jet and surfacc temperatures, the temperature of the surrounding gas, via

entrainment, also affects heat uansfer rates at the surface. Several authors including Goldstein et a1. (1990) and Hollwonh and Wilson (1984) defined correction factors in

terms of adiabatic surface temperature for h (or Nu) calculated using AT=(T.-Tj). The

intent is to quantify entrainment effects on impingement heat transfer. As the definition

of the adiabatic surface temperature by each author differs slightly, such correction

factors may be confusing. The adiabatic surface temperature is also a function of all of

the flow and geometric parameters that affect impingement heat uansfer. Moreover.

measuring or predicting the adiabatic surface temperature disuibution in actual

applications is difficult. Therefore, its use is not very practical.

For close jet-to-nozzle spacings, i.e. <8d, the nozzle exit temperature is generally the

preferred T,. For an isothermal impingement surface, with this choice of T,, local h

profiles then in fact become profiles of heat flux divided by a constant. Hence, local h - or in non-dimensional form "Nu" -does not account for changes in local AT that may

result due to effects on the boundary layer of various flow and gwmemc variables.

Page 6: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING

Potential Core

Free Jet Region -- Impingement Region

Figure 1. n o w ficld of an impinging jet

These effects need to be quantificd in the range of geometric and flow variables for

practical applications.

A shon revicw of the flow smcture of impinging jets would be beneficial to

understand the discussions that follow.

Row field of an impinging jet may be dividcd into t h m characteristic regions:

Free jet

Stagnation flow - Wall jet

Depending on nozzle shape, its characteristic dimension and nozzlc-to-surfacc distance.

h e fra jet may display a potcntial core, a developing flow and a developed flow

regions. Figure I.

The potential core is characterized by a constant jet centerline velocity nearly equal

to the nozzle cxit velocity. The length of the potential core rrgion is determined by the

rate of gmwth of the mixing layer at the jet boundary. For a contoured jet nozzle,

Page 7: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

because i t produces nearly a flat velocity profile and low nozzle exit turbulence, the

potential core length is significantly larger than any other nozzle types (6-8d).

D o s d o p (1969) and O b t (1980). However, for a particular nozzle type, the actual

length of the potential core may dccrcase as the turbulence level at the nozzle exit

increases, Saad ct al. (1992). For a commonly used nozzle type, i.e. square-edged

orifice. Hollwonh and Wilson (1984) repon a third length which agrees with those

reponed by O b t for an orifice of similar design. The% flow measurements were

mostly done at low temperatures. The centerline velocity of heated impinging jets in

industrial applications decays faster, hence resulting in shoner potential core lengths.

Kataoka et al. (1984).

In thedeveloping flow region. axial velocity decays as the jet spreads. Eventually, a

bell shaped profile is approached which can be described by a Gaussian disuibution,

Manin (1977). I t has been shown by several investigators that the turbulence level

continues to increase beyond the potential core region in the developing and developed

free jet regions. For closely spaced multiple jets, the turbulence intensity at the

centerline increases more rapidly and to higher values as compared with those for single

jets. Saad el a1.(1992).

The nozzle-to-jet spacing at which the maximum stagnation point heat transfer

occurs relates to the velocity and turbulence development at the centerline of the jet.

Heat transfer under jets emerging from contoured nozzles displays a maximum when

the impingement surface is at 6-8d away from nozzles. Gardon and Akfirat (1965)

proposed that the maximum occurred at the location of maximum centerline turbulence

The data by various researchers, more recently by Kataoka et al. (1987). confirms this.

Hence, the location of the maximum heat uansfer may be closer to the nozzle exit, or

may not exist at all, if the turbulence level at the nozzle exit is already high.

In the stagnation region, flow makes a W rum. Here, static pressure first increases

sharply with the corresponding drop in axial velocity, then drops as the flow accelerates

along the impingement surface. Hence, in this region, there is significant favorable

pressure gradient on the surface. Studies by Gutmark et al. (1978) and Saad (1981)

with slot jets consistently reported that effect of stagnation on free jet mean flow is not

felt bcyond 0.2H from the impingement surface, and on axial turbulence velocity

beyond O.OSH which is also in agreement with the results of Obot (1980) for circular

jets.

Page 8: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING 1153

The end of stagnation region in the lateral direction, defined as the location where

pressure gradient becomes rcro. is reported to be 0.35-0.5H from the impingement

point, Schaucr and Eustis (1963). Gardon and Akfirat (1965). Kumada and Mabuchi

(1970). Cadek (1968) and Saad ct al. (lW2). Beyond the stagnation region. in the wall

jet region, the pressure gradient in the lateral flow d i c t i o n is essentially rcro while the

fluid boundary layer over the impingement surface grows. Of the two sides of the wall

jet boundary layer, the impingement surface side shows typical effects of a

conventional boundary layer while the outer region has features of a free turbulent jet.

For a confined jet, if the confinement and impingement surfaces are sufficiently long,

the wall jet boundary layer grows to reach the confinement surface, thereby enclosing a

recirculating flow.

For a multiple jet system, another region of interest is the location where the wall

jets fmm adjacent jets meet. Characteristics of this region, as expected, highly depend

on the type of outflow used. This is a region of high turbulence; therefore, heat transfer

is enhanced, Saad et al. (1992). For slot impinging jets, if symmemcal exhaust pons in

the confinement surface are available, Saad et al. (1992) and Polat and Douglas (1990).

then the spent flow is directed upwards in the mid point between the jets without

causing cross flow effects on adjacent jets. Saad et al. reports that when the flow aspect

ratio, S N , is greater than 1.5, individual jets in such a system behave like independent

single jets, i.e. non-interacting.

Effects that are present in the industrial systems of impinging jets, such as surface

motion, exhaust port location, cross flow, entrainment and surface through flow

significantly change the boundary layer development over the impingement surface.

Moreover, due to the same effects, local temperature driving force for heat transfer is

also modified. Consequently, heat transfer rates at industrial impingement surfaces

reflect the combined effects of change in local shear rates and temperature driving

force.

Local Profiles

It is frequently said that the local dismbution of impingement uansfer rates has little

engineering value because a moving impingement surface, typical of industrial systems,

automatically integrates the local profiles. This is generally m e . However, for cases

Page 9: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

e 30 q i m o ~ l aqi j o pua aqi si3aUal em!u!m al(L 'aueld aizzou aql WOIJ ieme mg ueqi

ssai asepns e uo sa%u!dm! la1 asualnq1nl le!i!u! mole aiaqm maisB e ~ o j @s!dB are

ain%!cl j o salgo~d ~ajsuen ieaq [em1 aql u! em!xem pue em!u!m uo!leu%eis-jjo a u

.sa[zzou iuaserp aqi uaamiaq aJepns luwrauguo~ aq1 U! L lpx iyaumis

qe iuads Bu!lsneqxa i q paieu!m!ja s! maisk iacald!l[nm s!qlloj moU ssou

JO i m g a a q ~ .asepns luama%u!dm! aql mo l j ieme ms palem1 pue ap!m m u 01 are

salzzou 'maisis ~ajBu@u!dm! s!q~ 10s '(0661) selsnoa pue lelod 'eue uado 191 %OZ

q1!m s ~ a l ald!ilnm Bu!iswnu! alojaraqi p m d s L laso l~ l o j an salgo~d E un%!d 'aJepns

~uama%u!dm! aql m o l j ieme m q z paieml an ' m u OZ=M ' S ~ Z Z O U i a ~ h u a ~!id![la

a u '(~1661) .@ la ielod 'eare uado lac g s lnoqe q~!m maisis ral aId!ilnm Bug3clalu!

-uou 'paseds ilap!m BJO iallols al%u!s e lapun S! n~ [em1 j o salgo~d z arnl!d

.~ueuodm! ,ban amosaq L e u salgo~d [em[ 'siaclu!lu!dm! j o saxoj

s!meuiporpLq aqi xapun uo!ieuojap 01 alq!idassns s! aseps luama%u!dm! aql aIaqm

Page 10: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING 1155

purely laminar boundary layer from its minimum thickness at the stagnation point. The

increase from the minimum to the secondary maximum is due to the enhanced uansport

characteristics of a boundary layer in transition to turbulence. Beyond thc off-

stagnation maximum, fhc Nu profiles decline again with growing fhickncss of the

turbulent boundary layer.

4 I 4- - - -7- r/L R J L d L

Although the jet-to-surface spacing is still less than 8w, because of the very close

jct-to-jet spacing, these off-stagnation minima and maxima disappear in the profile

under the jets with 20% open area. Figure 3. This is typical of a syslcm where jet-to-jet

interaction is significant. Saad ct al. (1992).

60

20

H / v . 5 Re - S I H = 0.5 - E l 0 0 ..... 1 2000 - 20600 --- 25800

- ,**--.. - / - - - - I - Id---C' -- '- I \----- CZC-- ._ -.

8 0 , , ' - \.--*- 0--- 0' x-

.- /' ----a'

... k. , .--0-*

0

-

-7 .5 - 5 . 0 -2.5 0.0 2.5 5.0 7.5

Distance f r o m niddle J e t Centerllne, y l w

Figure 3. h f i l e s of local Nusselt number for interacting multiplc impinging slot

jets

Page 11: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

1156 POL AT

Distance t ram Nozzle Centerline, y / w

Figure 4. Profiles of surface p n s s u n relative to nozzle exit pressure for a single

slot jet atH=2.5w

Average Coeficieno Average Nusselt number. G, correlation for the jet system shown in Figure 2 is

- Nu =0.0314 Re0.76

valid in the ranges 3.2 5S/HS.4 and 16.000<Re<58.000.

And for the jet system shown in Figure 3 is

- Nu =0.094 ~ e ~ . ~ ~

valid for 8,00O<ReQ6,000.

A comparison of the average heat transfer rates on the basis of equal fan energy

requires that the jet Reynolds number for the f=3% system be 6.7 times higher than that

for f=ZO% system, provided nozzle discharge coefficients are equal. Within the range

of applicabilily of these correlations. Re=58.000 and 8.700 satisfy this requiremcnr

Page 12: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING 1157

0 1 I 10 5 0 5 10

Dlrlance from Stagnation Polnt. r/d

700

li r 600 - C

500 U - -

400 V L w -

300 u

+ ; 200 w I - 0 100- " 0 A

Figure 5. Pmfiles of local Nusselt number for a single round impinging jet: Effect

of jet-trnsurface distance, Gardon and Akfirat(1965)

0 - 6 J m m

- PI= 28.000 AT. ZO'C

- .

- - -

Using the above correlations, the correspondingG values are 83 and 44 for the f=3% and 20% systems respectively. Economic considerations would dictate the use of the

3% open area system because of higher heat wnsfer rates. However, relatively uniform

local profiles under the 20% open area system can be of interest if the product is

susceptible to deformation under high pressure gradients at the surface for the 3% open

area system. Figure 4.

Another option for providing more uniform, but lower, heat transfer at the surface is

to increase nozzle-to-surlace distance. As depicted in Figure 5 for round jets, when

H%-8w, the secondary p e a s disappear. A more bell-shaped heat transfer profile is

%en at the surface. In parallel, forces applied on the product due lo deceleration and

acceleration of the flow alsn decrease.

Jer Flow

Like any boundary layer flow, for a given set of geometric conditions, impingement

heat mnsfer rates increase as jet flow increases. m e dependency of an averdge Nusselt

Page 13: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

Figure 6a. Correlation of impinging slot jct data as a function of flow aspect ratio.

S/H. Sad(1981)

1.0

0.8

0 .6

0.4

0.2 - Nu

~ e " ( H / W ) ~

0 .10

0.08

0.06

0 . 0 4

0.02

number on Reynolds number is expressed as Nu = b Rea. Depending on the study, the

- - 0.33 ! SIH ! 1.33 1.51 SIH ! 4 - - 8 S H I V ! 2 4 8 ! H l v ! 2 4 - 3.3301Rc129.160 f<;:,&! 20,740 - - -O.ZIS(SIH~. - n . " ( ~ l v ) ~

. - - =

1 RP65(H,v)-0.80

- - /' - 0 .63 ( s I H ) ~ ~ ~ - - - I - I -

I I I I 1 1 1 1 1 I I I I , I

0.2 0 . 4 0 .6 0.8 1 0 2.0 4 .0 6.0

proportionality constant " b andthe exponent "a" are reported as functions of geometric

parameters. Ww. SiH. Slw or I. In thc stagnation region, when H <8w (or 8d), it is

generally observed that "a" is close to 0.5, a value typical of laminar boundary layer flows. With increasing averaging distance from stagnation point, i.e. Slw, for reasons

explained earlicr, Re exponent becomes higher, indicating a boundary layer in transition to turbulent flow.

Flow Cell Proportion. S/H

For single slot jets, or non-interacting multiplc jets, i.e. S/H>1.5, and for 8 W w

524. Saad (1981) found that the exponent "a" was constant for all practical purposes at

a value 0.65, Figure 6a. This agrees well with the value 0.67 reported by Martin (1977)

for 2cWw<80. As S/H decreases, i.e. the jets become more closely spaced, the turbulence created in the rcgion where the wall jets from the adjacent jets approach

each other affects heal Wnsfer. Saad repons an increase in the Re exponent from 0.65

to 0.8 almost linearly as S N decreased from 1.5 to 0.375. Figure 6b.

Page 14: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

f low Ce l l Proporl lon. S/H

Figure 6b. Dependency of avenge Nuswll number on Reynolds number for slot

jets

Figure 6c. Dependency of average Nusselt number on Ww for slot jcls

. 1.0

-0.8

-0.6-

m - 0 .4 -

-0.2

0

0 0

8 ! Hlv ! 24 - m

F " = c ( H / v ) -

- / - PC

0 5,500 - - 0 11.000

15.000

I I I 1 1 1 1 1 I I I I I

0.3 0.4 0.5 1 .O 1.5 2.0 3.0 4.0 5.0

Flow C e l l Proportion, S/H

Page 15: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

Obot et al. (1980) reports an extensive list of design correlations available for round

jets. The Re cxponent in the correlations range from as low as 0.33 to as high as 1.06.

However. Obot et al. note anomaly of such low and high Re exponents. The majority

of studies reported values ranging from 0.5 to 0.8.

Nozzle Geomerry

I t is generally recognized that the nozzle design appreciably affects the

impingement surface heat and mass transfer profiles. Different nozzle designs produce

different nozzle exit velocity and turbulence profiles. Moreover. the nozzle exit

turbulence is also affected by the nozzle design. As free stream turbulence enhances

heat transfer in a laminar boundary layer, it is then expected that the nozzle geometry

effect be more imponant for H<8w (or d).

A majority of laboratory studies have used contoured entry nozzles, Cadek (1968).

van Heiningen (1982). Saad et al. (1992). Polat et al. (1991a). Although this type of

nozzle is impractical for industrial use, for study of the effects on impingement

transport phenomena of other design variables, and also for computer simulation

studies, i t provides uniform nozzle exit conditions, i.e. flat velocity and turbulence profiles with low nozzle exit turbulence.

Using contoured entry slot nozzles. Saad et al. varied the nozzle width from 3 mm to

13.3 mm. A consistent increase in turbulence intensity at the nozzle exit plane. I, from

0.65% for the narrowest nozzle to 0.8% for the widest one was measured. perhaps

specific to his equipment. With increasing distance from nozzle exit, these small differences in I were shown to grow into big differences. Consquently, at H=8w, Nuo

was 17% higher for the widest nozzle than that for the narrowest nozzle.

Hardisty and Can (1980) studied heat transfer characteristics of impinging slot jets using 8 different typcs of nozzles in thc range 3.000<Re<13.000. They measured the

discharge coefficients of the nozzles, Figure 7, and found that the centerline Nu from

different nozzles can bc correlated using the effective slot width (w'=Cow) as the

characteristic dimension instead of the nozzle width, w. It is unlikely that the use of w'

would completely eliminate variation in Nu due to the nozzlc design. This variation is

partially due to the turbulence effects, discussed above. Some researchen tried to

incorporate a turbulence enhancement factor to their correlations. Kataoka et al. (1987).

Page 16: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING 1161

1.0 - NOZZIC wldth

D v.3 m m (conrtsnt) U

; 0 9 - c Nozzle u - shape Sqrnbol

a, 0

li U . a y 0.7 - 1J a = i r . ?2 ', 7 r 0 6 0.6 - \I A

-- 0

0.5 , 7

0 I I I I I , , I I I 1 I L I I I I I I ,

I 2 s 5 7 1 0 ~ 2 3 s 7 t 0 5 z 3 s

Reynolds number. Re

Figure 7. Discharge coefficients for various type slot nozzles. Hardisty and Can(1980)

For indusmal applications, however, such correlations are not practical: turbulence is

not an easily measurable or prrdictable quantity.

Consistent with the theory, both Hardisry and Can, and Saad (1980) repon that, for

geometrically similar nozzles, narrower nozzles gave higher heat transfer coefficients.

Obot (1980) has compared heat transfer and discharge coefficients for round jets

issuing fmm various types of nozzles and orifices. He also repons higher average heat

transfer coefficients for jets emerging from sharp entry nozzles when H<8d.

Differences between the uansfer coefficients from different types of nozzles tend to

decrease with increasing H, when H>8d.

Jet-lo-Surface Separalion

The effect on impingement hear transfer of the jet-to-surface distance is again

related to the flow and turbulence characteristics of a free jet. Figure 5 shows the

general characteristics of local profiles at various distances away from the nozzle exit

for a round jet. For especially close spacings, because of the influence of the nozzle

Page 17: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

Extent or Heat Transfer Surrace. S/H 4 3.2 2 1.33 t too I I I

Figure 8. Effect of nozzle-to-surface distance on average Nusselt number for slot

jets

: 80 i (U n

5 Z

= 60 a '9 ," > Z a rn a

40 > a

20

geometry and size on velocity and turbulence development, the effect of jet-to-surface

distance on Nu could vary depending on the nozzle design.

P d ~ t el a t . ( 1 9911) Slnqlr Jet v a n t l t l n i n p t n (1982) Sinpl. Jet

A Cobk(1968)SlnplrJ1t - 0 $806 ( 198 1 ) M u l l l p l e Jet1

-

S I W . 8 ( r . 6 . z ~ ~ ) - Re .21,000

I I I I I I I I I 0 2 4 6 8

For contoured entry slot nozzles in a multiple jet system. Saad repons that average

Nu in the range 4w<H<8w is almost independent of Ww. Figure 8. The trend

displayed by average Nu for single jets from contoured slot nozzles. Polat ct al.

(1991a,b), van Heiningen (1982), Cadek (1968). is to increase with increasing distance

from the nozzle exit for Ww values of up to 6w. In contradiction to bath of h e w

uends, not shown in Figure 8, Wedel's (1980) data for sharp entry multiple slot nozzles

show a continuous decrease with increasing Ww.

Nozzle-to-Surface Spaclng, H/w

For non-interacting impinging jets, for H/w>8. Saad's analysis indicates that

logarilhniic dependency of average heat transfer on Ww is a linear decrease with slope

-0.8. For interacting jets, the linear dependency on Ww is still valid: however, the

Page 18: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING 1163

slope becomes less sensitive to Ww as S/H ratio becomes smaller, Figure 6c. This is

opposite of the trend that is reported by Manin (1977) for arrays of round nozzles.

Journeaux ct al. (1992) repons that for round jets for 1dUdc4 and 35.000<Rc

<I 17,003, average Nu is almost independent of Wd. This agrees well with thc data of

Wedel for multiplc round jcts. For 2dUd<l2 and 15,WO<Re<60,000. Obot (1980)

found a steady decrease in average Nu with increasing distance from the nozzlc cxit.

Hc reports a value of -0.2 for the (Hld) exponent.

Jet-to-Jet Separation

The effect ofjet-to-jet separation distance, 2S/w, is naturally coupled with other

effects that exist in a multiple jet system. In a confined mulliplc jet system with spent

flow exhausting from one cnd of the channel that is formed between confinement and

impingement surfaces. cross flow effects exist. Although there is less concern for cross

flow, a discussion on unconfined impingement systems is not useful because, for

thermal efficiency reasons, they are not commonly used in commercial systems.

The studies by Saad et al. (1992) and Polat and Douglas (1990), and Polar et al.

(1991a.b) isolated the effects ofjet-to-jet separation from the cross flow effects by

providing symmetrical exhaust ports in between slot nozzles at the confinement surface.

Saad et al. showed that this system can be thought of as an assembly of repeated flow

cells with an aspect ratio of S/H. They classified the multiple jets as "non-interacting"

when the internozzle spacing, 2S, is sufficiently wide and "interacting" when the

internozzle spacing is not wide enough. Their data indicated that for 4dUw<24 when

S/H>1.5, a multiple impinging jet system without cross flow effects is effectively an

assembly of single jets, therefore single jet transfer data can be used just as effectively.

For S/H<1.5, because the hcat uansfer profiles changed significantly due to jet-to-jet

interactions, special multiple jct data is needed. This is also evident from Manin's

correlation results displayed in Figure 9. Polat and Douglas (1990) reported heat

uansfer results for the Ww and S/H combination. 5 and 0.5, that Saad (1981) predicted

as being the combination that would give the highest heat transfer for a system of slot

impinging jets without cross flow effects, Figure 9.

T-T Sriegl and Diller (1984a.b) proposed to use an entrainment factor, F = 2 . 1 0 TI-T,

predict heat transfer in a multiple jet system from the single jet data. This method was

successful only for widely spaced jas , indicating that interactions in a closely spaced

Page 19: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

FIOV c e l l Ratio, S I H

Figure 9. Effect of flow cell ratio on average Nusselt number as a function of

nozzle-to-surface distance. Hlw: Comparison of the data and predictions

jet system affect both the temperature and flow fields. Journeaux et al. (1992) applied

the approach to the case of unconfined round jets and correlated their results in terms of

F.

A heat balance for the control volume shown in Figure 10, gives the following

As indicated by this relation. F increases with decreasing percent jet open area, f.

i.e. with increasing jet-to-jet separation distance. Hence, as SW increases for a fixed

HJw. the F value for h e cell increases indicating that the spent fluid temperature,T,, is

approaching toT,. Figure 1 I compares average Nu under round impinging jets

predicted using correlations by Martin (1977) for a single round jet and arrays of round

jets, and by Journeaux et al. for a row of round jets using F values of =0.0.3 and 0.5. , The Manin's single jet correlation results agree well with the results of Journeaux et al.

at F=O, i.e. for a case where the nozzle exit temperature is the same as that of the

environment. Figure I I results also show that when the isothermal jet results of

Journeaux et al. are corrected for enuainmenl in a multiple jet system, using F as a

Page 20: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING

Figure 10. Control column for heat balance in a flow cell

.......... Mnnin11977). S indc .......... Manin (1977). Multiple c I

Jovmcalu ecd. (1-2) 0 t, 120- ,, ----- Jomeaux e~a1. (1992) 09 .......... Joumuur =I a1.0992) 03

E, roo- z - - g 80

- YI ..-. 5 60 -

40 4 I 1.0 2.0 3.0 4.0 5.0

Flow Cell Ratio. SM

Figure 11. Effect of flow cell ratio on average heat transfer for round jets:

Entrainment factor, F

comction factor, they agree very well with the predictions for multiple jets. It is

interesting to note that, as expected, the multiple jet results approach the F 4 . 3 line at

small SM values and the F 4 . S line at higher S/H values.

Spenr (oi Cross) Flow Effects

In a confined jet system, without symmetrical exhaust of spent flow between jet

nozzles, cross flow effects on heat transfer should be considered. In such systems, the

Page 21: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

flow from the intermediatc jets has to cross the jet flow from nozzles closer to the

exhaust pons. As cwler spent flow from the intermediate jets accumulates and flows

across other jet flows, significant changes in heat transfer is expected to occur as both

local shear rates and temperatures are modificd by this superimposed flow.

Saad (1981) measured local and average heat transfer profiles under impinging slot

jets with and without the effects of cross flow. Hc noted a 15% to 30% decrease in

average Nu when cross flow was only I to 2 times the jet flow. This decrease was

found to be insensitive to Re and f in the ranges 5.700<ReQ0,7W and 3<1<8.

Saad et al. (1980b) reported local and average Nu numbers for staggered anays of circular jets for 2<Syld<4, 3<Sx/d<6 and 3,35O<RcQ1,500. They found that, in the

range I d / d < 3 , the cross flow effect does not significantly affect heat aansfer for up to 3-5 jet rows. In industrial systems, exhaust pons are typically provided at every 3 to 10

rows of jets. Hencc thc effects may be greater for larger number of rows in a group.

The correlation based on a one-dimensional model presented by Galant and Martinez

(1982) for cross flow effects may then be used to predict the extent of such effects.

Jet Temperature and Hwridiry

In impingement drying, the variation in fluid properties from nozzle exit to the

surface is substantial and thus must be considered. In their study of superhcated steam

drying of paper. Bond el al. (1990) used Manin's correlation for anays of round jets to

calculate impingement heat transfer coefficients in their system. With fluid properties evaluated at a reference temperature given by the one-third mle suggested by Chow and Chung (1983),T, = fi+ :T.. and transpiration effects included using the film theory correction for heat transfer, given laier, they obwincd good agreement between the

predictions and their data, Figure 12.

Richards and Florschuetz (1986) measured impingement heat transfer coefficients

under conditions of varying humidity at the jet nozzles. Their primary objective was to

evaluate existing mcthods to calculate viscosity and thermal conductivity of airlwaar

vapor mixtures. Their results indicate that neglecting the effect of humidity does not

produce large errors. < 10% for humidity ratios up to 0.25 (kg waterkg dry air).

Evaporation at rhe Surfoce

In drying, especially during the constant rate period, evaporation rate at the surface

is high. The reported impingement heat and mass transfer coefficients do not include

Page 22: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING 1167

Temperature. Deg. C

,025 r

E \ rn .02 Y

a- " ,015

rn C .- 2 & .o 1

D a N .-

,005

E D

0

Figure 12. Impingement drying using superheated steam: Comparison between the

data and Manin's cornlalion with and without transpiration corrections

,' , - Bond el al. (1990) - Corrected lor Transpilalion /' - ,

No1 Correcled lor Transpiralion 4

#' 4'

#'

evaporation effects. Therefore, these coefficients should k corrected appropriately

when used i n the design o f dryers. Bird. Steward and Lightfoot (1960) describe several

approaches to determine the form o f the comction factors. Here, only the comction

factors based on the f i lm theory, which wen: successfuUy applied to impingement flows

by Crotogino and Allenger (1979) and Bond et al. (1990). are given.

0 : , 100 , / ZOO 300 400 SO0

Heat transfer coefficient with evaporation at the surface:

Mass transfer coefficient with evaporation at the surface:

Surfnce Morion

Impingement drying of continuous sheets of material involves impingement to take

place on a rapidly moving surface, a feature which may change heat and mass transfer

characteristics substantially.

Page 23: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

Rej -35400 H/v -2 .5

100

-16 - 1 2 - 8 - 4 0 4 8 I Z 16

Otstance from Nozzle Centerllne. y /w

Figure 13. Profiles of local Nusselt number for a single slor jet impinging on a

moving surface: Effect of surface motion

Polar and Douglas (1990) and Polat et al. (1991b) provided the local heat transfer

profiles at a rapidly moving surface under confined slot jets for interacting multiple jets

and a single jet respectively. They expressed surface motion non dimensionally as the surface-to-jet mass velocity ratio, MvS. For the jet Reynolds numbers in the range

18.000-35,400, Mvs was varied from 0.029 to 0.34 by varying the surface speed from

0.5 ro 9 m/s. Figure 13 profiles at HJw=2.5 and Re=35,400 indicate that for this close

jet-lo-surface spacing, the largest effect on the local hear transfer is felt in the wall jet

region of the side where surface motion is towards the nozzle centerline. This was

perhaps due to the dominating effect of reduction in local AT near the surface when

cooler temperature spent fluid is dragged by the surface motion into the jet region. 7he

net effect of surface motion on average hear transfer rates was found to decrease heat

transfer. For both interacting and non-interacting jets, this reduction was correlated only as a function of MvS as ( I + M V ~ ) - ~ . ~ .

Page 24: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING 1169

Journeaux et al. (1992) more recently reponed a similar analysis for an array of

confined and unconfined round jets impinging on a rapidly moving surface. Their conclusion was that the impingement heat transfer was not appreciably affected for Mvs

values up to 0.6, i.c. twice the maximum value that was obtained by Polat et al. for slot

jets. Insensitivity to swiacc motion in h e case of round jets may be explained with the

fact that, in this case, the spent flow has a higher degree of freedom to spread on the

surface.

Surface Throughflow

At a permeable impingement surface, such as paper and textiles, convective

transpon rates can be enhanced funher by withdrawing some of the jet flow through the

surface. Baines and Keffer (1976. 1979) measured the effect of through flow on local

shear stress, and by analogy reponed enhancement in heat transfer due to through flow.

Because analogy between momentum and heat transfer does not hold for the stagnation

region of impinging flows, validity of those results is questionable. Saad (1981) and

Obot (1982) measured increase in impingement heat transfer for a limited set of

conditions for slot and round jets, thus, they did not repon any correlations. However.

both noted a uniform, linear increase in local heat transfer profiles with through flow.

Enhancement in local impingement heat transfer due to through flow was measured

for multiple. Polar and Douglas (1990), as well as a single impinging jet, Polat et al.

(l991a.b) using a permeable heat flux sensor, Polat et al. (1990). They expressed through flow non-dimensionally as the through flow-to-jet mass velocity ratio, MuS.

Figure 14 shows enhancement of local heat transfer profiles with increasing MuS for the

single jet case. As reponed by Saad and Obot, enhancement is nearly uniform

everywhere in the profiles.

Polar el al. noted that, based on a heat balance near the surface, enhancement due to

through flow is best expressed in terms of Stanton number. They found that increase in

Stanton number due to through flow is proportional only to the through flow parameter Mus with a proportionality constant of about 0.17 for both interacting and non-

interacting jets. Their maximum Mus value was 0.023. These results indicate that the proponionality constant may be valid even in a much wider range of geometric and

flow parameters.

Referring to Figure 14, the significance of the surface through flow effect for industrial applications is apparent that by using Mu, = 0.0121. heat transfer is nearly

Page 25: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

0 1 -16 - 1 2 - 8 - 4 0 4 8 I2 16

Distance from Nozzle Centerline, y/w

Figure 14. Profiles of local Nusselt number for a single slot jet impinging on a

permeable surface: Effect of surface through flow

doubled for heat transfer surfaces of any half-width. S, over the broad range 0.8 - 6.4.

In this range. f=25%-3%. the through flow rate is only 4.8% to 38.7% of the jet flow

rate. Adding to this impressive enhancement of convective heat transfer at the surface.

there is funher enhancement inside the permeable smcture due to intimate contact with

gas, Polat. 0. (1989). With such impressive enhancement features, the combined

impingement and through flow drying c e d n l y opens possibilities for even low

permeability media.

Because impingement heat and mass transfer has attracted the attention of many

researchers, there are quite a few correlations available in literature for this type of

flow. Obot et a1.(1980). The majority are limited to a narrow range of flow and

geometric variables. The Figure 6 correlation by Saad (1980) for multiple slot jet

systems without cross flow effects, and the following correlations by Manin (1977) are

therefore suggested.

Page 26: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING

where f. = (60 + ~ ( W Z W - ~ ) ~ ) ~ ~ ~ .

Range of applicability is

I ,500<Re<40,000 0.008<f<2.Sf0

2<Ww<80

Note that the cornlation by Martin for slot jet systems was developed using data for

a system where spcnt flow was allowed along the length of the nozzles. This outflow

arrangement natl~rally imposed three-dimensional effects on heat transfer rates.

Therefore the predictions using Martin correlation on Figure 9 are lower than the

cxpcrimental data for slot jets where exit pons were provided symmetrically on both

sides of the jets. In the range that is most relevant for industrial applications, however.

Martin's correlation agrees reasonably well with the data.

Although Saad's correlation is valid in the range 8<Ww<24, he observed that for

4 W w < 8 , his results were relatively insensitive to Ww showing only about 5% increase with decreasing H.

For arrays of round jets, again the following correlation by Martin is suggested

Range of validity

102,000<Re<100.000

O.W4<f<0.04

2cH/d<12

Correction factors that are suggested for the effects of surface motion, evaporation.

and/or through flow in the above sections can be applied to these Nu correlations when

such effects are present.

Due to the complex interaction on impingement heat transfer of turbulence, mean

flow and temperature fields, modified due to the effects of flow and geometric

Page 27: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

variables, these correlations should still be used cautiously. Once the conceptual dryer

design is completed, it is common practice to check the validity of the results using a

small scale dryer.

P E W CONSIDERATIONS

Economic reasons dictate maximizing the heat transfer per unit fan energy per unit

heat mnsfer area. Assuming the jet nozzles m the major resislance to flow, h e

following relation between the fan energy, nozzle discharge coefficient and the nozzle

exit velocity is valid.

Using this relation, an optimal spatial arrangement for jet nozzles can be determined

while keeping one of the characteristic lengths, i.e. w. H or S, constant. Manin (1977).

using his comlations for impingement heat transfer, gives the following optimal values

on the basis of a constant H.

Staggered Array of Round Array of lets Slot lets

0.015 0.072

It is interesting to nore that the optimal S/H value for both types of nozzles is the

same and very close to rhe value that Saad et al. (1992) reported to be a critical value,

S/H=0.75, where jet-to-jet interactions start affecting hear transfer at the stagnation

point.

One scenario for the design procedure would bc to choose the values of H and d to

achieve the optimal Wd value while keeping the maintenance and operation constraints

for a particular application in mind. The S value is thus fixed by the optimal S/H and H

values. With the selection of the nozzle type and nozzle exit velocity - which

determines the total volume of gas for the system - the dryer area is found from the total

drying requirements of the system. The drycr design is completed with the layour and

sizing of the duct work.

Page 28: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING 1173

The nozzle exit velocity could be limited if the product is susceptible to deformation

under the hydrodynamic force by the jets. In this case, slot jets should be preferred

over round jeu. This is because, for the same blower rating, slot jets give about the

same heat transfer as the round jets at lower jet velocities. On the other hand, with slot

jets, the total volume of gas recirculated in the system is almost three times greater.

CONCLUSIONS

High drying rates achieved with impinging jets make this technology attractive for a

number of industrial applications. Sensitivity of heat and mass transfer rates, however,

in such applications to a number of flow and geometrical parameten makes the design

engineer's job difficult as to the selection of a design basis and control strategy. Here a

concise review ofthe imponant effects on impingement transport phenomena is given. Although the dixussions are based on impingement heat transfer, as analogy to mass

transfer is well established, they equally apply to both.

Remember that, because of the interactive nature of a number of effects, laboratory

investigations directed to quantify certain effects may sometimes produce results and

conclusions that are not directly relevant to indusmal use. On the other hand, because

the numbcr of effects is high, an experimental program to fully characterize and

develop general correlations for impingement flows would be an impossible task.

Correlations by Manin (1977) for arrays of slot and round jets when moderate cross

flow effecu are present, and by Saad (1980) for arrays of slot jets without cross flow

effects are given here because of the wider range of variables they covered.

The same Nu correlations can be used to predict impingement transfer coefficients if

surface motion and through flow effects are important with the multiplication or

addition of appropriate facton. For slot jets, as suggested by Polat and Douglas (1990) and Polat et al. (1991b). multiplying average Nu with the factor ( I + M V , ) - ~ . ~ would

account for the surface motion effecu. For round jets this effect is not important even

for quite high values of Mv,. Joumeaux et al. (1992) . For surface through flow effects,

average Nu number under both slot and round jets can be modified by adding the term

0.17 Mv, Re R, Polat and Douglas (1990) and Polar et a1.(1991a). These correction

factors can be used either alone or in combination depending on the casc.

Based on the discussions provided here on general mechanisms of effects, design

engineers should usc their best judgement to account for other effects specific to a

particular case.

Page 29: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

NOMENCLATURE

CD : nozzle discharge coefficient

d :jet nozzle diameter

f : fraction open area (T-T,) F : enuainment factor. 1 (Tj-TA

: heat transfer coefficient

: nozzle-ro-surface distance

: volumeuic gas flow rate per unit heat transfer area

:radial direction

: jet-to-jet separation distance in flow direction for a two dimensional jet m y

: jet-to-jet separation distance in lateral direction for a two dimensional jet m y

: jet-to-jet separation half distance : nozzle exit temperature

:Reference temperature

: surface temperature

: nozzle exit velocity

: nozzle exit velocity for slot jet

: nozzle width

: fan energy

Non-dimensional numbers

Ww : nozzle-to-surface distance

Re : Reynolds number, = P V,=i(or w) P

Mv, : surface motion parameter, surface-to-jet mass velocity ratio

Mu, : surface through flow parameter, through flow-to-jet mass velocity

ratio h d(or w)

Nu : Nusselt number,= - k

h : Prandtl number SiH : flow cell ratio

S/w : averaging distance from jet centerline

w12S : fraction open area for slot jet systems

REFERENCES

Baines, W.D. and J.F. Keffer (1976), Shear Smss and Heat Transfer at a Stagnation Point, Int. Heat Mass Transfer, Vol. 19, pp.21-26.

Page 30: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

IMPINGEMENT DRYING 1175

Baincs. W.D. and J.F. Keffer (1979). Shear S a s s Measurements for an lmpinging Air Jct, Transactions of the Technical Section, Canadian Pulp and Paper Assoc. Vol. 5, pp. 39-44.

Bird. R.B., W.E. Steward and E.N. Lightfoot (1960). Transpon Phenomena. John Wiley & Sons. Inc., New York.

Bond. J.F., R.H. Crotogino, W.J.M. Douglas. A.S. Mujumdar, and A.R.P. van Heiningen (1990). Impingement Drying of Paper in Supcrhcated Steam in the Constant Rate Period. Presented in lntemational Drying Symposium. Prague.

Cadek. F.F. (1968). A Fundamental Investigation of Jet Impingement Heat Transfer, Ph.D. Thesis, University of Cincinnati.

Chow. L.C. and 1.N Chung (1983). Evaporation of Waar into a Laminar Stream of Air and Supcrhcalcd Slcam. Int. J Hcat Mass Transfcr. Vol. 26. No. 3. pp. 373-380.

Crotogino. R.H. and V. Allcngcr (1979). Mathematical Model of the Papridryer Process. Transaction of thc Technicvl Section. Canadian Pulp and Paper Association, Vol. 5, No. 4, pp. 84-91.

Dosdogru, G.A. (1969). Uber die Ausfuhrung von Schlitzdusen im Untershallbereich. Mitteilungen Heft 2, Forschungsgesell schaft Dnrckmachinen e.V.

Galant. S. and G. Manincz (1982). Cross Flow lnflucncc upon lmpingc~nent Convccltvc Hcat Transfer in Circular Arraysof Jcts. Pmcccdings of the 7th lnt. Heat Transfcr Confncncc, Munchcn, Vol. 3. pp. 343-347.

Gardon, R. and J.C. Akfirat (1965). The Role of Turbulence in Determining the Heat Transfer Characteristics of lmpinging Jets, Int. J. Hcat Mass Transfer. Vol. 8, pp. 1261- 1272.

Goldstein. R.J., K.A. Sobolik and W.S. Swl(1990). Effect of Enwinment on the Heat Transfer to a Heated Circular Air Jet lmpinging on a Flat Surface. J. Heat Transfer. Vol. 112, pp. 608-611.

Gutmark. E., M. Wolfshtein and I. W y g n a n s ~ (1978). The Plane Turbulent Impinging Jet. J. Fluid Mech.. Vol. 88. Pan 4. pp. 737-756.

Hardisty. H. and M. Can (1980). An Experimental investigation into the Effect of Changes in the Gwmcuy of a Slot Nozzle on the Heat Transfer Characteristics of an lmpinging Jet, Roc Instn Mech Engrs. Vol. 197C. pp. 7-15.

Ilollwonh. B R and S.I. Wilson (1984). Enminmcnt Effccls on implngemcnt Hcat Tnnsfcr. Pan l-Mcasurcmcnts of Hcatcd Jct Vclocity and Temperature Distributions and Hccovcry Tcmpcrdturcs on Tugct Surfdcc, J. Heal Transfer. Vol. 106, pp. 797-803.

Journeaux. I. R.H. Crotogino and W.J.M. Douglas (1992). lmpinging let Heat Transfer in Calender Contml Systems: Pan I and 11, in preparation.

Kntaoka. K.. H. Shundoh and H. Matsuo (1984). Convective Heat Transfer Between a Flat Platc and a Jct of Hot Gas lmping~ng on it. Drying'84, Ed. Mujumdar. A S.. McGraw Htll Book Co.. pp. 218-226

Kataoka. K.. R. Sahara. H. A w and T. Harada (1987). Role of Large-Scale Coherent Smcmrcs in lmpinging Jet Heat Transfer. J. Chem. Eng. Japan. Vol. 20, No. I, pp. 71-76,

Page 31: Heat and Mass Transfer in Impingement Drying

Dow

nloa

ded

By:

[200

7 K

ing

Mon

gkut

's U

nive

rsity

of T

echn

olog

y Th

onbu

ri] A

t: 10

:54

31 O

ctob

er 2

007

Kumada, M. and I. Mabuchi (1970), Studies on the Heat Transfer of lmpinging Jets. Bull. of JSME. Vol. 13. No. 55, pp. 77-85.

Manin, H. (1977). Heat and Mass Transfer Benuccn Impinging Gas Jets and Solid Surfaces. Advances in Heat Transfer. Academic Press, Vol. 13, pp. 1-66.

Obot. N.T. (1980). Flow and Heat Transfer for lmpinging Round Turbulent Jets. Ph.D. Thesis. Chem. Eng. Dept., McGill University.

Obot. N. T. (1982). Effect of Suction on lmpingement Heat Transfer, Proceedings of the 7th lnt. Heat Transfer Conference. Munchen. Vol. 3, pp. 389-394.

Obot. N.T.. A.S. Mujumdar and W.J.M. Douglas (1980). Deslgn Conelaiions for Hcat and Mass Transfer Under Various Turbulent lmplnging Jct Configurations. Drying'80. Ed. Mujumdu. A S . McGraw Hill Book Co . Vol. I , pp.388402.

Polat. 0 . (1989). Throughflow w i n g of Paper, Ph.D. Thesis, Mdjill Univenity.

Polat. S. and W.J.M. Douglas (1990). Heat Transfer Under Multiple Slot Jets lmpinging on a Permeable Moving Surface, AIChE Journal, Vol. 36, No. 9, pp. 1370-1378.

Polai. S.. A S. Mujumdu and W.J.M. Douglas (l99la). Impingemeni Hcat Transfer Undcr a Confined Slot Jct. Pan I. Effccl of Sufiacc Through flow. CJChR. Vol. 69. pp. 266-274.

Polnt, S., A.S. Mujumdarand W.J.M. Douglas (1991b). Impingement Heat Transfer Under a Confined Slot Jet, Pan 11: Effects of Surface Motion and Through flow. CJChE, Vol. 69, pp. 274- 280.

Polat. S.. A.R.P. van Heiningen and W.J.M. Douglas (1990). Sensor for Transient Heat Flux at a Surface With Through flow. Int. J. Heat Mass Transfer. Vol. 34. No. 6, pp. 1515-1523.

Richards, D.R. and L.W. Florschuetz (1986). Forced Convection Heat Transfer to AirIWater Vapor Mixtures, Proceedings of the Eighth Lnt. Heat Transfer Conference. San Fransisco, Vol. 3, pp. 1053-1058.

Saad. N.R. (1981). Flow and Heat Transfer for Multiple Turbulcni lmpinging Slot Jets. Ph.D. Thesis. Chcm. Eng. Dcpr.. McGill Univenity.

Saad. N.R., S. Polat and W.J.M. Douglas (1992). Confined Multiple lmpinging Slot Jets Without Cross flow Effects, Int. J. Heat and Fluid Flow, Vol. 13. No.1, pp. 2-14.

Schaucr. J.J. and R.H. Eusds (1963). 7hc Flow Development and Heal 'Transfer Characensucs of Planc Turbulent lmpinging JCLS. Techn~cal Rcpon No. 3. Mech. Eng. Dcpl.. Stanford U n i v c n ~ ~ y .

Striegl, S.A. and T.E. Diller (1984aa). The Effect of Entrainment Temperature on Jet lmpingement Hcat Transfer, I. Heat Transfer, Vol. 106, pp. 27-33.

Smcgl, S A. and T E Dlllcr (1984b. An Analys~s of ihc Effect of Envalnmcnl T c m p c n t u ~ on Jct Imp~ngcmcnl lieat Transfer. J Iicat Transfcr. Vol 106, pp 804. 810

van Heiningen, A.R.P. (1982), Heat Transfer Under an lmpinging Slot Jet. Ph.D. Thesis. McGill Univenity.

Wedel, G.L. (1980), Air Impingement Hcat Transfer, Tappi, Vol. 63, No. 8, pp. 89-92. I