オホーツク海の夏季海況に就いて 梶浦欣二郎
緒言
オホーツク海の海況に就いては,襯測資料が少い爲め,現在迄余り研究されていない.こ
の報告は・主に昭和17年夏水路部と中央氣象毫とが協同し,中央氣象喜搬測船凌風丸によつて行
われた観測に基き,この海域全般の海況を明らかにしやうとしたものである.
- 幅:2159px
- 高さ:2745px
- ファイルサイズ:295.7KB
§1 海水の層重欺態
オホーツク海内部及び太平洋の千島近海に於ける代表貼の温戯曲線を第1圖に示す.之れによれば,オホーツク海内部表暦水は陸水の流入,冬期生成された海氷の融解等のために低鹽(塩)分を呈し,主として日射のため夏期は相當に昇温している.75m乃至100m層を中心として水温が著しく低く,所謂中冷層が存在するが,これは宇田博士により輪ぜられた如く,冬期表層水が冷却され,又結氷に際して生じた高密度海水が封流により沈降したものである.中冷層と表層の聞には安定度の非常に大きい不蓮続層がみられるが,中冷層と下層との境界は明瞭でなく,150m層前後に於ては安定度が中性に近く・鉛直封流が割合に起り易いことを示している.200m以深になると水温,鹽分共ほぼ一様に増加し,800m乃至1000m層で水温が極大となる.この原因は後述する如く,深層に於ける太平洋水の流入によるものと思われる.千島列島聞の海峡はすべて水深1500m以下であるため,それ以深は3000m以上の底迄ほとんど水温の変化がみられす,1.8度乃至1・9度を示し,オホーツク海固有の水塊が存在するものと考えられる.
- 幅:1659px
- 高さ:2139px
- ファイルサイズ:129.5KB
§2水温,鹽分の分布
a)水温,鹽分の水干分布
表面に於ける鹽分双ぴ,年変化を考慮して補正を加えた水温の分布を書いたものが,第2圖である.表層では,すで指摘されている様に,中部千島近海に低温,高鹽分海水が存在する.
これはオホーツク海水が太平洋に流出するに際し,海峡附近で渦動か発達し,海水の鉛直混合が盛んに行われるためと思われる.中部樺太沖に見られる低温,高鹽分水は,鉛直断面に於ける水温,鹽分の分布を併せ考えると,地形性の湧昇流によるものらしい.南樺太東方海面に広がつている32%以下という低鹽分海水は,厚さ10m乃至25mに及び,その起源は明らかでないが,今迄の観測資料によれば.日本海側にはかかる低鹽分水は見られず,又陸水の影響と考えるには余り広範園に存在し過ぎる様である.あるいは融水により生じた低鹽分水が,風のためこの海域に集積したものかも知れぬ.北樺太の近海も著しく低鹽分(30%以下)であるが,これは予想される通り,黒龍江排水の大部分其の他シベリアの陸水が多量に流入する爲めであらう.カムチヤツカ半島南部近海には周囲より低温,高鹽分水がみられるが,之は観測当時東風が卓越していたため,海岸近くに湧昇流か起り,鉛直渦動が発達したためではなからうか.宗谷海峡より高温,高鹽分の■島海流末枝が流入し,南千島近海に迄及んでいるが,10m層になると,北海道北沖約50浬附近に冷水があらはれ,下層水の上昇が考えられる.25m以深になると,水温,鹽分の分布は表暦とやや趣きを異にし,中冷層のそれに近くなる.一般に表層に於ては顯著なる海流は認められず,中部千島附近よリオホーツク海側の冷水の太牛洋側に流出している事が目立つのみである.
中冷層に於ける水温分布を知るため,最低水温の分布図を書いたものが第3図である.これによると,北千島附近よ1りは太平洋側の温暖水が流入し,中部千島附近よりは冷水が太平洋に流出している・又新知島北方の高温なるは150°E附近を北上する暖流末伎の影響であり,南千島,北海道北沖の高温なのは宗谷暖流の存在を示すものと考えられる.最低温層の深度は,54°E以北50m,以南は千島沿海を除き大■75mで,一搬に北より南に向うにつれ深くなつているが,特にカムチヤツカ半島の南端附近では深さ125mとなり,冷水が太平洋より流入する暖水の下に潜入している有様がうかがえる・塩分は一般に周辺部に多く,52°N,149°Eを中心として低塩分海域(約0.3%過少)が存在する・深さ20Omに至る迄はほぼ同様な水温,塩分の分布傾向を示すが,カムチヤツカ南端附近で沈降した低温・低塩分水は深くなるにつれ次第に北千島北方一帯に広がる.300m深前後では150°Eを境として,樺太側がカムチヤツカ側より約0.5℃低水温となり,塩分も少く,特に樺太側の南部が低温,低塩分である.この傾向は深層に迄あらはれ,500m以深になると水温,塩分の変化は少いが,それでも南部は北部よりやや低温,低塩分を示し中部千島附近にそれが著しい・800m乃至1000m層に至ると,北千島附近の海峡より温暖なる太平洋水が流入し,等深線に平行に北西に流動する様に見える.そのため中及南千島の北方のみ周園より約0.1℃低温,0.1%低塩分となつている・1200m層に到ると水温変化はほとんどなくなつている. ここに注目すぺきは,200mより1000mの深層に至る迄ボ一ソリ海峡の南方は周園の太平洋水より低温,低塩分(約0.3℃低温,0.2%低塩分)を示し,上層とは逆にオホーツク海の冷水がこの海峡を通して太平洋に流出していることである.
b)水温,塩分の鉛直断面に於ける分布
一般に深さ25m前後に顯著なる水温躍層が存在し,50m乃至150mの間は0℃以下の低温を示している.然し第4図よりも知れる様に,南部カムチヤツカ西沖では,この躍層が明瞭を欠き,表暦は鉛直混合の盛んなことが判る.千島沿海を除いては,一般に南西部で等温,等塩分線が最も降下しているが,これは深層流か海底地形に從つたために現れたものか,或はこの附近は,冬期生成される氷原の縁■に当たり,最も■流が発達するため,深層迄その影響が及んでいるのか,どちらかであらう.千島沿海では,すでに水平分布の項で述べた如く,中部千島近海の200m乃至300m以浅に於ける海水の鉛直渦動による混合,北千島近海に於ける太平洋水の流入と,その下層への冷水の沈降がみとめられる.叉宗谷海峡の底層よりは,水温2℃乃至3℃の日本海系水が流入し,100m乃至150m層を北海道北岸にごく接近して南東に流れ,北海道東端145°E附近にて二枝に分れ,一枝は南干島北沖に達し,一枝は北に向を■じて中知床岬の東方海面に達する.又亜■湾内には水温0℃以下の冷水が10m層附近に存在し,暖水と冷水とは,南樺太と北海道の間の海域で復雑に交錯している.新知島北方の約400m以深で等温,等塩分線共に降下していることは第5図より知れる通りであるが,この原因としては、100m暦前後に於ける暖水の流入のため,北方より流来したオホーツク海の冷水がその下に沈降すること,及び深層に於けるオホーツク海水がボーソリ海峡より流出する際に,一部が新知島にさえぎられて沈降する事が考えられる.中部千島東沖では上昇流のみられることはすでに述べたが,これは等温,等塩分線がそのあたりで凸出していることより知り得る.叉水平分布図にも現はれているが,北部オホーツク海56°N,144°E附近に,周園と全く異り表面より150mの海底に至る迄,水温1・5℃前後,塩分:3.9%前後でほとんど変化のない海水が存在する.この様な特異な海水が常にこの附近に存在することは,他の時期に行われた観測資料よりも確認し得る.その成因としては,附近に海堆が存在するため,海水の水平流動が相当ある時には,ここに渦動が発達し海水の鉛直混合が充分行はれる事が考えられる.千島の諸海峡を通して,その両側の海水の混合は,200m以深ではあまり顯著ではない.千島列島に沿う太平洋側の鉛直断面図をみると,北海道南方約150浬附近に,高温,高塩分なる黒潮末枝が明瞭に認められ,又暖流の一分枝は得撫島南方150°E附近にもみられる.寒流は南千島,北海道沿海,上述の二暖流分枝の中間,及び中部干島の南方の三枝が存在する.然し太平洋側寒流域にみられる中冷層は,オホーツク海内のもの程低温ではなく,水温0℃乃至1℃で100m深前後に存在する.太平洋側に於ける上述の結果は宇田博士の研究と大旨一致している.
- 幅:1753px
- 高さ:2153px
- ファイルサイズ:209.8KB
- 幅:1666px
- 高さ:2447px
- ファイルサイズ:229.5KB
- 幅:3145px
- 高さ:2219px
- ファイルサイズ:276.7KB
- 幅:2999px
- 高さ:3943px
- ファイルサイズ:441.8KB
§3 海水の流動
a) 力學的地形図 400m層を基準とし,深さ100m即ち中冷層に於ける海水の流動を表わしたものが第6図である.800m層を基準とした図もこれと大旨同様な傾向を示すが,流速はやや過大となる.図には0.01ダイナミツクメーター毎に等量線を引いてある故,50°Nに於いて,その間隔が180浬あるとき1■/秒の流速を示す.一般にオホーツク海内部は非常に密度流が微弱なることが知れる.南西オホーツク海に於いて,海水は中知床岬の沖合より等深線にほぼ平行して東北東に流れ,北千島附近より西流した海水と152°E,49°N近傍で衝突し,一部は南下して中部千島の諸海峡より太平洋に流出し,残りは北に■向し,50°Nより53°N迄の間はカムヤツカ半島のごく沿海を除き,一般に北流となる.このため北方より海岸沿いに南下したる海水は流路を■じ北樺太の東方海面に環流を生ずるに至る.太平洋側の流れはすでに§2にて述べた通りになつているが,中部千島より流出したる海水は,カムチヤツカ東岸を流下した海水と共に直ちに南下を始め,南千島沿海を流れる寒流とは直接関係を持たない様に見える.
b) σtの水平分布 今100m以深の海は平行ソレノイド場であると仮定して海水流動を推察すると,200m深のあたり迄は力學的地形図によるものと大旨同様な流動を示すが,それ以深になると,ボーソリ海峡より流出するオホ一ツク海水以外に千島列島の両側の海水の蓮絡はほとんどなくなる.然し深さ1000mに至ると再び僅かながら北千島附近より太平洋水の流入を認め得る.
c) 力學的断面図 鉛直断面に於ける比容の分布よりみれば,10m乃至25m深に存する密度不連続層はほぼ水平に位置し,流動の緩慢なることを示している.
d)計算流速 密度流としては,表層より100m深附近迄はほとんど等速のところが多く,毎秒数■以下である.然るに,海面に於いては海岸近くではあるが0.2節程度の流れが観測されている.
これより考えるとオホーツク海表層の海流の原因は風が主となつているらしい.
e)等密度面上の塩分分布 イセントロピツク面(Isentropic Surface)上の塩分の分布を調べれば、種々の海水の水平混合の有様が理解される。第7図はσt=26.00 の面を示しているが,この面は表面乃至50mに位置する.これによるとオホーツク海内の塩分は32.5%前後,太平洋側は32.9%以上で,両海水の混合状態はすでに§2にて述べたことを裏書きしている.
叉亜庭湾内の陸水系低塩分水南千島近海の日本海系暖水,中部樺太東沖の湧昇流による高塩分水のみられることも,既述の通りである.
ボーソリ海峡の南方に暖流系水が接近しているが,300mより深くなると,逆にこの海峡よリオホーツク海の低塩分水が太干洋に■散している.一般に200m以深では太平洋側がオホーツク海側より高鰹分で,余り水平混合を起していない.
- 幅:1673px
- 高さ:2698px
- ファイルサイズ:249.4KB
- 幅:1600px
- 高さ:2219px
- ファイルサイズ:158.9KB
§4 水色及び透明度の分布
第7図は,水色,透明度の分布を示し,今迄の研究と大旨一致している.この海域の大部分は水色4であり,水色番号の大きいのは,すべて鉛直渦動が発達していると見なされる海である。又透明度は大旨水色と相伴って変化している。47°N〜48°N,145°E〜151°Eの範園に存在する周囲より特に清澄な海水は,例年見られるらしいが成因は明らかでない.又54°N附近にも清澄な海水が存在する.
結語 極く海岸近くの海水流動については,観測■の数が少い故確かなことは解らないが,オホーツク海の夏期一般海況はやや明瞭になつたと思われる.この研究では,今迄考えられていた東樺太海流の存在は認められず,又左旋の大環流なるものも考えられなかつた.然し,千島近海に於ける海水の流動については,ある程度既往の研究と一致し,上暦200m深附近迄は海峡の両側の海水は互に蓮絡しているが,それ以深になるとボーソリ海峡を通つて流出するオホーツク海水を除き,ほとんど相互の流動はみられない.千島沿海には又沈降流等も存在し今後の研究に待つ問題が多い.其の他,北千島より流入する太平洋水の消長は,オホーツク海東部の海況に大きな影響を及ぼすであらうから,158°E附近を北上する暖流分枝の消長とオホーツク海の海況の変動との關係等も研究か望まれる.
この研究を終るに当り,終始御指導を賜わつた日高教授に御禮申上げます.
[参考文献]
(1) 丸川久俊:漁業基本調査報告 第7冊ノ2、第8冊ノ1 大正5年,大正6年
(2) 重松良一:「オホツク」海の海流及親潮の起源等に就いて「水路要報」第12年第8號昭8年
(3) 宇田道隆:東北海域に於ける中冷層の分布,成因,蓮動に就て「海と空」第15巻第12號昭
10年
(4) 神尾秀二:千島列島浩海温度表に就て「海と塞」第11巻第5號昭6年
(5) 第4,第5回日本海北部海洋齪測報告「海洋時報」第6巻第1號
(6) 宗谷海峡表面航走翻測報告「函館海洋気象台,海洋時報」第2號昭22年
(7) 宇田道隆:日本海、黄海、オホーツク海の平年各月海況「水試報告」第5號昭9年
(8) 宇田道隆:昭和8年盛夏に於ける北太平洋の海況「水試報告」第6號昭12年
(9) 宇田道隆:東北海域に於ける海況の変動に就て「水試報告」第9號昭12年
(10) 宇田道隆:日本近海に於ける水色透明度「海と室」第10巻第8號昭5年
(11) K. Suda:On the Dissipation of Energy in the Density Currnt(2nd Paper)
中央気象台■文彙報 第10雀第2號,1936
(7’) 昭和12年5,6月施行 水路部 軍艦「駒橋」の観測結果:未獲表
- 幅:1619px
- 高さ:2472px
- ファイルサイズ:281.7KB
On the Hydrography of the OKhotsk Sea in Summer. (Abstract) By K.Kajiura.
In this paper,the general oceanoglaphic conditions of the Okhotsk Sea were investigated on the basis of obserations made in summer of 1942 on board the ship「Ryofu-maru」, which was conductcd by the Hydropraphic Department in cooperation with the Central Meteological Observatory.
It was found that a conspicuous discoontinuity of temperature exists at the depth between 10m and 25m, and the mininmm tenperature of about -1.5℃ is at the depth of about 100m.
Also, an examimtion was made with regard to the state of flowing of water in and out of this region, and it was found that the Pacific water flows into this area through the North Kurile and the Okhotsk water flows out through the Central Kurile.
It was also found that, within this area,the primary current is generally weak, and the current drifted by wind prevails.
The Velocity Distribution of Wind Currents in the Eastern Part of the Equatorial Pacific. By K.KAJICTRA (Geophysical Institute, Tokyo University)
Abstract
Wind driven currents in a baroclinic ocean are discussed following the ideas of SVERDRUP and REID, and the current velocities in the upper layers of the eastern equatorial Pacific are calculated on the assumption that the inertial forces and the lateral friction are negligible. According to the present theory, currents may be devided into two parts, of which one is the pure drift currents and the other is the relative currents, and it is found that the results are in good agreement with the present knowledge of the currents near the equator. In these regions, the relative currents are much stronger than the pure drift currents and the north-south component of the velocity is very small compared with the east-west component.
Introduction
Recent theories on the wind driven ocean circulation have shown the circulation in the upper layers of the oceans being the result chiefly of the stresses exerted by winds. The stream lines of oceanic mass transport derived from solutions of a vertically integrated vorticity equation account for many of the gross features of the general circulation (MUNK,1950).
Mass transport methods, however, add only little to the knowledge of the vertical structure of currents. In this respect, HIDAKA (1950) treated this problem by using the velocity components themselves instead of the mass transport, but the results were not satisfactory. Different approach to this problem is attempted here which may be successful in a region where lateral stresses of the currents can be neglected.
According to the ideas of SVERDRUP (1947)and REID(1948),the pressure field in a baroclinic ocean can be completed by the knowledges of the wind stresses and the vertical structure of water-mass. Therefore the problem to be solved is reduced to find currents under the influence of wind stresses in the known pressure field. This problem is similar to that discussed by EKMAN(1906).
Theory
In a region where lateral stresses and field accelerations can be neglected, the equations of stationary motion of water can be written;
式(1)
Here u and v are the x and v components of velocity where x- and y- axes are directed toward the east and north, respectively, with the origin on the equator, and the z- axis is taken to be positive downwards. And \psi\ is the latitude, taken positive to the north of the equator,\omega\, the angular velocity of the Earth’s rotation,\rho\, the density,p, the pressure and \mu\, the eddy viscosity.
Writing
[Fig.107_021_02]
where i^{2}=-1, the equations(1) are reduced to,
式(2)
In this equation \alpha\ includes \rho\ and \mu\ both of which must be functions of z.
The variation of the density with depth, however, is too small to modify the results and the knowledges of the eddy viscosity are not enough at present to draw reliable conclusions as to the variation of a with depth. Therefore, on the assumption ofαto be constant in the vertical direction, the general solution of the equation (2) is easily obtained as follows
式(3)
where A and B are arbitrary constants.
Introducing the boundary conditions such that [Fig.107_021_05] at the surface and W=O at a sufficiently great depth, in which [Fig.107_021_06] and \tau_{x}\ and \tau_{y}\ are x and y components of the wind stresses, W can be determined definitely when F is known.
In a baroclinic ocean it is reasonable to assume the horizontal velocity and the horizontal pressure gradient vanish at a moderate depth below the sea surface. If d is the selected depth of no motlon, the vertically integrated equations of motion from the surface to the depth,d, on the assumption that [Fig.107_021_07] at the depth,d, are
式(4)
K.Kajiura:The Velocity Distribution of Wind Currents.
where [Fig.107_022_01]
The equation of continuity is,
式(5)
By cross-differentiation of the two equations of (4) with the aid of (5), it follows that
式(6)
where \tau_{x}\ and\tau_{y}\ are assumed to be independent of x,and \psi\is taken to be y/R, in which R is the radius of the earth.
Substitution of (6) into (5) leads to
式(7)
Here the continental boundary is assumed at x=x0 where no mass transport is present across the boundary.
By insertion of both (6) and (7) into (4)the following equations are obtained,
式(8)
Assuming the Reid’s model of the vertical structure of mass, such that
[Fig.107_022_06]
where [Fig.107_022_07] is the elevation of the sea surface, the pressure at any depth may be determined by a integration of the hydrostatic equation.
From the condition that the pressure gradient vanishes below a moderate depth, it follows that,
式(9)
when the depth of no motion is assumed to be sufficiently great compared with the depth of the upper homogeneous layer.
By use of the above equations (9),the following relations are obtained,
式(10)
and from the equations(8)and(10),it follows that
式(11)
where C is a constant of integration.
The pressure distributions in
式(12)
Similar expressions hold for [Fig.107_023_02]
If the equations (8), (9), (10), (11) and (12) are combined together, the pressure gradient at any depth can be computed from the wind stresses alone.
Formulae for the current velocities in the upper layers.
The arbitrary constants A and B in the equation (3) can be easily determined from the boundary conditions as the horizontal pressure gradient does not vary with depth in the upper homogeneous layer and it vanishes at the depth, d.
For the sake of simplicity, the Lower boundary is replaced from z=d to z=\infty\.
This simplification may not greatly modify the result.
The current velocities in the upper layers are given by,
式(13)
In the equations 13) the first term of the right hand side represents the pure drift current similar to that discussed by Ekman and the second term represents the relative current which depends on the distribution of density.
The calculation of currellt veloeities.
The density distribution being related to the stress gradient, it is desirable to determine the accurate distribution of wind stresses, but it is impossible at present because of the lack of available observations. Therefore, we assume the sinusoidal stress distribution in the equatorial region of the North Pacific similar to the one assuiizecl by Reid (198),such that
[Fig.107_023_04]
and
[Fig.107_023_05]
where
[Fig.107_024_01]
This formula for \tau_{y}\ is applied only to the north of 4°N and it is assumed that \tau_{y}\ decreases linearly from 4° N to the equator and it vanishes in the south latitude.
Numerical values used in calculation are as follows;
[Fig.107_024_02]
Fig.1 is the vector representation of the wind stress and Fig.2 shows the variation of currents with latitude. The drift current, depending on the absolute value of the wind stresses, varies its intensities according to the value of \tau_{y}\, while the relative current which is mainly determined by the gradient of \tau_{x}\ in the north-south direction is only little affected by the presence of \tau_{y}\ except
near the coastal region.
Fig.3shows the velocity distribution in the upPer layers. The drift current decreases exponentially with depth and the depth of frictional influence depends on the latitude and the value of eddy viscosity, so that the currents at the depth deeper than about 100m are al most negligible except in the very vicinity to the equator, when \mu\ is assumed to be 10^{2} C.G.S. In the surface layer, the drift current seems to make narrower the counter current and the maximum velocities of the counter current appear in the sub-surface layer.
The maximum calculated velocity of the counter current is about 20cm/sec which is only half of the expected value of the velocity of the counter current.
This discrepancy may be ascribed to the under-estimate of the wind stress in this region, or the neglect of the thermodynamic effect. On the whole, relative currents prevail compared with the drift currents. This conclusion is, however, not always reasonable as the drift current is developed by the temporary wind but the relative current is maintained by the stationary wind stress and is influenced only little by the temporary wind. Just under the equator, the above theory fails and other factors such as lateral stresses, field accerelations and the thermodynamic processes must be taken into consideration.
Divergence and convergence of currents on the surface can be computed from the following equation;
[Fig.107_026_01]
The calculation shows that no remarkable regions of divergence or convergence exist, except in the very vicinity of the equator.
Concluding remarks.
In the region discussed in this paper, the neglect of lateral stresses and inertia terms may not introduce serious result, because the horizontal stresses are rather small compared with the vertical stresses, even when the lateral eddy viscosity is taken to be 10^{3} c.g.s. except along the boundary line between the equatorial current and the counter current, where both stresses may become equal in magnitude.
When the drift current and the relative current are co-existing, the stationary state cannot be reached in a baroclinic ocean without some mechanism which re-establishes the distribution of density of water disturbed by drift currents.
Lateral mixing processes and thermodynamic processes may be the factors which take part in this mechanism.1n the present paper, however, the assumption of homogeneous water in the upper layer diminishes these difficulties and simplifies the treatment of phenomena.
Further investigations about the model of a water-mass are desirable but it seems very difficult to treat it exactly. The formation of water-mass is chiefly the result of thermodynamic processes such as evaporation and precipitation.If the thermodynamic processes act uniformly over the large area without any dynamic disturbances, a water-mass is formed in which the density of water may become a function of the depth only. When wind begins to blow under this circumstance and when the unequality of the wind velocity exist, the surface water is accumulated somewhere, result:in the cause of the horizontal pressure gradient.This pressure gradient, reaching to the very bottom of the ocean, may produce the movement of the water in deep layers which changes the distribution of density. Thus the horizontal pressure gradient caused by the surface elevation will be cancelled soon by the re-distribution of density, and when the stationary state is reached, the water in the deeper layers will be at rest. In these ways the inner pressure gradient which becomes negligible at a moderate depth, may be generated.
The present paper is only a first step to study this interesting problem of the wind current in a baroclinic ocean, and many objections may be raised against the looseness of the treatrnents. Nevertheless we believe that the density distribution in the ocean may play an important role for the distribution of velocity of the wind current.
Acknowledgement.
The author wishes to express his sincere thanks to Prof. K. Hidaka for his support and encouragement during this work.
References.
Ekman, V. W.(1906):Beitrage zur Theorie der Meeresstromungen. Ann.d. Hydro.u.Marit. Meteorol.,34, pp.566-583.
Hidaka. K.(1950):Drift Currents in an Enclosed Ocean, Part 1. Gpophysical Notes(Geophysical Institute, Tokyo Univ.).3, No.23.
Munk, W. H.(1950):On the Wind-Drived Ocean Circulation・J. Meteorol.,7, PP・79-93.
Reid, R.0.(1948a):The Equatorial Currents of the Eastern Pacific as Maintained by the Stress of the Wind.J.Marine Res.,7, PP.74-99.
Reid, R.O.(1948b):A Model of the Vertical Structure of Mass in Equatorial Wind-Driven Currents of a Baroclinic Ocean..」Marine Res.,7, PP.304-312.
Sverdrup, H. U.(1947):Wind-Driven Currents in a Baroclinic Ocean;with Application to the Equatorial Currents of the Eastern Pacific. Proc.Nat. Acad. Sci.,33,pp.318-326.
- 幅:2219px
- 高さ:460px
- ファイルサイズ:31.8KB
- 幅:2113px
- 高さ:200px
- ファイルサイズ:21KB
- 幅:1766px
- 高さ:207px
- ファイルサイズ:12.3KB
- 幅:2373px
- 高さ:393px
- ファイルサイズ:34.7KB
- 幅:467px
- 高さ:220px
- ファイルサイズ:6KB
- 幅:406px
- 高さ:120px
- ファイルサイズ:3.7KB
- 幅:493px
- 高さ:174px
- ファイルサイズ:6.4KB
- 幅:2099px
- 高さ:387px
- ファイルサイズ:26.8KB
- 幅:1746px
- 高さ:253px
- ファイルサイズ:18.5KB
- 幅:2047px
- 高さ:233px
- ファイルサイズ:14.5KB
- 幅:2013px
- 高さ:247px
- ファイルサイズ:16.5KB
- 幅:2360px
- 高さ:233px
- ファイルサイズ:22.9KB
- 幅:2346px
- 高さ:467px
- ファイルサイズ:38.7KB
- 幅:3066px
- 高さ:107px
- ファイルサイズ:18.6KB
- 幅:803px
- 高さ:112px
- ファイルサイズ:6.3KB
- 幅:2353px
- 高さ:233px
- ファイルサイズ:23.4KB
- 幅:2347px
- 高さ:240px
- ファイルサイズ:23.3KB
- 幅:2386px
- 高さ:234px
- ファイルサイズ:24.7KB
- 幅:2892px
- 高さ:673px
- ファイルサイズ:57.7KB
- 幅:233px
- 高さ:113px
- ファイルサイズ:3KB
- 幅:2945px
- 高さ:713px
- ファイルサイズ:80.4KB
- 幅:953px
- 高さ:114px
- ファイルサイズ:8.3KB
- 幅:859px
- 高さ:106px
- ファイルサイズ:7.2KB
- 幅:3072px
- 高さ:233px
- ファイルサイズ:47.6KB
- 幅:2660px
- 高さ:240px
- ファイルサイズ:41.6KB
- 幅:866px
- 高さ:1752px
- ファイルサイズ:51.6KB
- 幅:1953px
- 高さ:1912px
- ファイルサイズ:122.3KB
- 幅:1813px
- 高さ:1893px
- ファイルサイズ:105.8KB
- 幅:2233px
- 高さ:1399px
- ファイルサイズ:119.6KB
- 幅:1673px
- 高さ:220px
- ファイルサイズ:20.8KB
The Effect of Winds on the Stratified Lake Water By K.KAJIURA (Geophysical Institute, Tokyo University)
Abstract
The currents and the accumulation produced by winds in the water consisting of two layers of different densities are investigated on the assumption that the eddy viscosity is negligible along the boundary of two water masses. The computed profile of the boundary surface is in food agreement with observation.
lntroduction
The currents and the accumulation of water in an enclosed sea and in a lake produced by the wind stresses have been investigated by many authors.
Most of them assumed, however,that the water is homogeneous in spite of the fact that the actual ocean and lakes are generally in stable stratification and there exist remarkable discontinuities of density below the surface in summer. In the homogeneous water, the horizontal pressure gradient due to the surface elevation caused by the wind reaches to the very bottom, and the whole water makes an circulation, en the other hand, in the non-homogeneous water, the movement of water changes the inner distribution of density and the inclination of isopleths gradually compensates the pressure gradient exerted by the surface deformation. Therefore in the stationary state, the features of the water movement in the homogeneous water and in the non-homogeneous water may be quite different. Moreover, the eddy viscosity is generally expected to decrease greatly owing to the increase of stability at the boundary of two water masses of different density. A. Defant (1932) and H. Arakawa (1935) discussed this problem, but they made an approximation of small amplitude of disturbances along the boundary of two water masses which may not be accepted when the elevation or the depression of the boundary surface becomes large. E. Palmen (1941) investigated the effect of the density stratification upon the elevation of the sea surface and concluded that when the considerably large area is concerned, this effect may be neglected because it needs very long time for the water movement to reach a stationary state.
In a small lake, however, the density stratification modifies the pattern of the vertical circulation of the water as the stationary state is easily attained.
Theoretical Consideration
We consider the motion of water in a canal consisting of two layers of different density. The equations of motion in a stationary state is,
式(1)
where the x-axis is the horizontal co-ordinate lying on the undisturbed surface and perpendicular to the coast with the origin at the coast and the z-axis is directed upward. u is the velocity along the.the x-axis, \rho\ is the density, \mu\ is the eddy viscosity, p is the pressure, and g is the acceleration due to gravity.
As equations of continuity we take
式(2)
where \zeta\ is the surface elevation, h is the depth to the bourldary of different water masses, and d is the depth to the bottom. Suffixes 1 and 2 are used to represent the quantities in the upper and the lower layer, respectively.
The boundary conditions are
式(3a,b)
where T is the constant wind stress exerted on the surface of water. When \mu_{2}\ is assumed to be neglegible along the internal boundary, the upper water feels no friction at the boundary surface, and the condition (3b) is reduced to
式(3c)
In a stationary state the water in the lower layer is at rest because the frictional effect of the upper water is interrupted at the boundary and the pressure due to the surface elevation is adjusted by the depression of the boundary surface. This condition is satisfied by the following relation,
式(4)
where H is the mean thickness of the upper layer.
The first equation of (1) is integrated from the depth of the boundary to the surface, taking the second equation into consideration,
[Fig.107_030_01]
and with the equations (3a) and (4) it follows;
式(5)
This equation can be integrated easily and results in,
式(6)
where [Fig.107_030_04] and A is a constant of integration. The positive sign in the second term of the right hand side is taken, because the elevation must be positive at some points.
The constant A is to be determined from the conservation of mass such that
[Fig.107_030_05]
where l is the width of the canal.
The slope of the surface is given by,
式(7)
At the point x=-A/b where no elevation of water exists, the slope is the same as that in the homogeneous water of the depth H, and to the up-wind side of this point, it is steeper and to the lee-wind side it is smoother.
From the equation (1) it foflows that
式(8)
where B and C are constants of integration.
The equations (3a),(4) and (5) yield,
B=h,
and the equation (2) becomes,
[Fig.107_030_08]
Assuming\zeta\ << h・we obtain [Fig.107_030_09]
Therefore we have,
式(9)
which shows that the vertical distribution of velocity is similar to thet in the
homogeneous water of depth h with smooth bottom.
- 幅:1893px
- 高さ:460px
- ファイルサイズ:25.7KB
- 幅:2146px
- 高さ:226px
- ファイルサイズ:17KB
- 幅:1807px
- 高さ:419px
- ファイルサイズ:26.5KB
- 幅:1827px
- 高さ:207px
- ファイルサイズ:12.3KB
- 幅:1746px
- 高さ:454px
- ファイルサイズ:24.4KB
- 幅:1186px
- 高さ:234px
- ファイルサイズ:15.7KB
- 幅:2966px
- 高さ:413px
- ファイルサイズ:30.6KB
- 幅:1973px
- 高さ:220px
- ファイルサイズ:15.4KB
- 幅:633px
- 高さ:207px
- ファイルサイズ:9KB
- 幅:447px
- 高さ:213px
- ファイルサイズ:5.5KB
- 幅:1842px
- 高さ:242px
- ファイルサイズ:17.2KB
- 幅:1893px
- 高さ:234px
- ファイルサイズ:17.3KB
- 幅:1193px
- 高さ:246px
- ファイルサイズ:18.5KB
- 幅:1945px
- 高さ:220px
- ファイルサイズ:16.9KB
The comparison with the observation
The above theory is applied to Lake Suwa, which is about 6〜7 meters deep and 4 km wide. According to J. Masuzawa (1950),there exists in this lake a thermocline in summer at the depth of about 5 meters and this stratification is not destroyed by the mixing caused by the prevailing winds of about 5〜10 m/sec which blows regularly in every afternoon. He found the very small eddy duffusion coefficient (about O.02 c.g.s.) near the boundary layer of different water masses, and this fact supports the assumption used in the present theory such as \mu_{2}\=0 at the boundary.
Fig.1 shows the observed distribution of temperature after Masuzawa in the morning and in the afternoon. In both cases the isotherms are almost coincident near the lee-wind coast, because the observations at St.46 and St.47 where carried soon after the wind set in and the stationary state was not reached yet. Fig.2 shows the calculated inclination of the boundary surface. The inclination of the isotherms of 22℃〜23℃ in the up-wind side in Fig.1 and that of Fig.2are in good agreement.
The numerical constants used in the computation are as follows:
[Fig.107_031_04]
- 幅:1725px
- 高さ:1266px
- ファイルサイズ:111.8KB
- 幅:1645px
- 高さ:706px
- ファイルサイズ:43KB
- 幅:2005px
- 高さ:273px
- ファイルサイズ:29.1KB
Conclusion
In a small lake the circulation of water due to stationary wind stress is greatly modified by the stratification of water of different density. The surface elevation of water is related to the thickness of the layer of lighter, upper water and the depth of the lake itself iS of little importance to the elevation, when it is sufficiently deep.
The slope of the surface is not constant but relate to the square root of the distance from some fixed point which is determined from various factors such as the width of the lake,l, depth of the upper layer, H, density difference, \alpha\, and the stress of the wind, T.
In a transient state, the movement of the lower water is chiefly governed by the hydrostatic pressure due to the surface elevation and the circulation pattern may be similar to the one in the homogeneous water.
The auther wishes to express his hearty thanks to Mr.,T. Masuzawa for his kind advice.
References
A.Defant:Beitrag zur theoretischen Limnologie.Beitr. Phys.freien Atmos.,19 143-150,(1932).
荒川秀俊:吹送流による湖水の循環 海と空15 437-444, (1935). (H.Arakawa:Circulation of Lake Water due to Wind Stresses).
E・Palmen:Zur Frage des Windstau in einem geschichtetem Meer. Soc.&Scient. Fenn.,Comm. Phys.一Math. XI Nr.7,1-11,(1941).
J.Masuzawa:On the stratification of lake water in summer, Lake Suwa, Nagano Pref.(in Japanese). Rep. Res. Inst. Nat. Res. No.15 (1950).
On the Influence of Bottom Topography on Ocean Currents. By Kinjiro KAJILTRA (Geophysical Institute, Tokyo University)
Abstract
Topographic effect on currents has been discussed by the method of small perturbation. An equation for the vertical motion is derived which includes a wide range of phenomena, and it is shown that the character of the deflection of currents by the bottom deformation is complicated since the different kinds of waves become dominant according to the scale of topography, and, moreover, since the horizontal shear of a current takes part in the phenomena. When the scale of the bottom deformation is very small, the stationary internal gravity waves prevail under a certain condition and when the scale is large, planetary waves develop in a horizontal plane.
1. Introduction
It is a well known fact that the distribution of density is considerably modified when an ocean current flows across a submarine ridge. Theoretical attacks to this problem, however, are very scarce owing to difficulties of mathematical analyses. In his very extensive studies on the character of ocean currents in homogeneous water, V. W. Ekman (1923) first examined the effect of bottom topography on slope currents, the velocity of which is constant with depth except in a layer influenced by the bottom friction. His conclusions are:The currents generally tend to follow the bottom contours in higher Iatitudes and in the northern hemisphere, they deviate to the right, passing over an elevation and to the left, over depressions of the bottom, and vice versa in the southern hemisphere. He extended his analysis later (1932) on the basis of hydrodynamic equations including inertial terms as well as frictional terms, but the major features were almost as before. Similar studies of the problem have been made by many meteorologists concerning to an air flow over a mountain.P・Queney (1947,1948) has developed an extensive study of perturbations of air flow and discussed the mountain effect, taking the vertical stability of the atmosphere into account. Recently, the influerlce of a large scale mountain range on westerlies has been discussed by many authors (J. Charney and A.Eliassen・1949・D・Colson,1949, B・Bolin,1950, etc.)on the basis of the vorticity equation. In these studies, frictional effects are neglected and the general flow is assumed to be zonal and uniform. According to the vorticity equation in a barotropic atmosphere, the horizontal divergence or convergence due to the effect of topography results in an increase or decrease of vorticity,respectively, and the deflection of stream lines can be expected on the downstream side which is followed by the planetary waves of Rossby type .
On the other hand, the effect of bottom topography on relative current in non-homogerleous water has been examined qualitatively by H. U. Sverdrup (1941). The conclusion of his analysis is that the observed deflections of currents, which seem to be maximum at the summit of the bottom ridge, may not be caused by bottom friction as was assumed by Ekman, but they are caused by modifications of the density distribution due to the effect of topography. In his theory, the assumption that the isopicnic surface must rise when a watermass approaches a ridge, includes essentially the existence of a vertical velocity as well as a horizontal velocity near the bottom, and this is inconsistent with the 「law of the parallel solenoids」 which is the fundamental concept in his discussion.
In these circumstances, V. B. Stockman (1947,1948) reviewed the former studies and discussed the topographic effect, using the concepts of mass transport and lateral frictions. He emphasized the prilmary importance of the distribution of wind system over the ocean which is the main catlse of ocean currents, and concluded that the deflection of currents by the bottorn topography may not be related to the Coriolis force, but is connected with the vorticity in the windfield, such that when a current crosses a sea-bottom elevation, the flow-lines deviate over its summit in the same direction as the vorticity in the wind field and just the opposite above the central part of a depression. Though his conclusions are very suggessive, his deduction of results seems to be very doubtful in respect to the procedure of simplifying the equations of motion, by means of
taking mean values of various quantities.
At any rate, it is a usual experience that the currents, flowing close to the coast or flowing slowly, are apt to follow the bottom contours. It is, however, not so exactly ascertained that in which direction they will turn when ocean currents pass over an elevation or depression of the bottom. Therefore, in the present paper, the effect of bottom topography on ocean currents has been reexamined on the basis of the hydrodynamic equations including inertial terms, taking the density distribution into consideration as far as possible.
2.Basic Equations of Motion
We assume that the field is stationary and may be divided into two parts, of which one is the basic field and the other is the field of small parturbations caused by irregularities of the bottom. In spite of the defects involved in the approximation of small perturbations, the qualitative features of a flow pattern may be estimated. We write the pressure, p=po+p1, the density \rho\= \rho\0十+\rho\1, the horizontal components of velocity u=uo十u1, v= v1 and the vertical component of velocity\omega\=\omega\1, where subscripts 0 and 1 refer to the quantities in the basic state and the disturbed state, respectively.
The basic current is assumed geostrophic and zonal defined by the following equations:
式(1)
and
式(2)
where the co-ordinate x, y and z are directed to the east, to the north and vertically upward, respectively, with the origin situated at the bottom, and λ=2(\omega\ sin (ψ) is the Coriolis parameter and、g is the acceleration due to gravity.
The equations of motion for stationary perturbations are:
式(3)
式(4)
and
式(5)
provided that the frictional and other small terms be neglected and that the basic current u0 is assumed to be independent of x.
When the fluid is incompressible and individual density is conservative, we have:
式(6)
and
式(7)
Cross differentiating (3) and (4) with respect to x and y, and taking (6)
and (7) into account, we obtain:
式(8)
and from (3) and (5),we have:
式(9)
Eqs. (8) and (9) may be simplified when adequate assumptions are introduced.
When the horizontal gradient of density is very small and the basic current maybe considered to be almost uniform , we have, from (8),neglecting the variation of velocities In the y-direction,
式(10)
and from (9),
式(11)
where U is the uniform basic velocity.
Here we shall transform v1 and \omega\1 into v’and \omega\’ by the following equations:
式(12)
From equations (1) and (2) ,we have,
[Fig.107_036_06]
where P_{0B} is the bottom pressure. The assumption of almost uniform basic current is satisfied when the condition [Fig.107_036_07] is fulfilled, in other words, when the baroclinic effect is neglected. Since the relative currents prevail in the ocean, this assumption can not be accepted. Our discussion, however, will be valid if the effect of the vertical shear of velocity does not play a dominant role. Thus, U may be given, regardless of (1) and (2), in our discussion.
Writting [Fig.107_037_01] both of which are assumed to be constant,we have from (10) and (11) , under the condition [Fig.107_037_02]
式(13)
and
式(14)
where [Fig.107_037_05] , and the term containing [Fig.107_037_06] is neglected,sinces, the coefficient of [Fig.107_037_07], is sufficiently small.The orders of magnitude of various parameters are tabulated in Table1.Here, we assume k_{\beta\} and ks to be positive, since we are chiefly concerned with steady and stable perturbations.
Table 1. Order of magnitude of various parameters.
[Fig.107_037_08]
In (14),we can see that when k >> k_{\beta\}, k_{\beta\^{2}} may be neglected and the equation will hold for any direction of a basic current.
- 幅:1914px
- 高さ:234px
- ファイルサイズ:13KB
- 幅:1900px
- 高さ:273px
- ファイルサイズ:14.3KB
- 幅:2480px
- 高さ:233px
- ファイルサイズ:22.6KB
- 幅:2500px
- 高さ:254px
- ファイルサイズ:20.4KB
- 幅:2527px
- 高さ:274px
- ファイルサイズ:20.7KB
- 幅:2200px
- 高さ:267px
- ファイルサイズ:17.8KB
- 幅:2187px
- 高さ:260px
- ファイルサイズ:19.4KB
- 幅:2659px
- 高さ:666px
- ファイルサイズ:87.1KB
- 幅:2619px
- 高さ:653px
- ファイルサイズ:87.3KB
- 幅:2153px
- 高さ:234px
- ファイルサイズ:22.4KB
- 幅:2746px
- 高さ:220px
- ファイルサイズ:34.1KB
- 幅:2333px
- 高さ:306px
- ファイルサイズ:22.9KB
- 幅:1513px
- 高さ:234px
- ファイルサイズ:15.8KB
- 幅:846px
- 高さ:246px
- ファイルサイズ:10.2KB
- 幅:927px
- 高さ:227px
- ファイルサイズ:12KB
- 幅:793px
- 高さ:247px
- ファイルサイズ:11.3KB
- 幅:1887px
- 高さ:247px
- ファイルサイズ:14.8KB
- 幅:2240px
- 高さ:253px
- ファイルサイズ:22.2KB
- 幅:938px
- 高さ:195px
- ファイルサイズ:12.8KB
- 幅:186px
- 高さ:206px
- ファイルサイズ:3.3KB
- 幅:160px
- 高さ:187px
- ファイルサイズ:3.4KB
- 幅:2373px
- 高さ:700px
- ファイルサイズ:72KB
3. Solution of the equations
A formal solution of (14) is,
式(15)
where [Fig.107_037_10] and A and」B are constants of integration.
Since p^{2} is positive when [Fig.107_037_11] and negative when [Fig.107_037_12], p is real or pure imaginary according to the value of k. Substituting (15) into(12),we hare,
式(16)
Here, boundary conditions for the vertical velocity icl are given by
式(17)
and
式(18)
where h is the height of the bottom deformation which is assumed to be independent of y corresporlding to the assumption that the velocity variation ln v-direction is negligible.
In the case when the bottom deformation is expressed by h=h0coskx and the mean depth is H, we have from (16) ,(17) and (18),as a particular solution:
式(19)
from (10) and (19),
式(20)
from (6),(19) and (20),
式(21)
and from (7),(19),(20) and (21),
式(22)
where [Fig.107_038_06]. ln equations(19) to (22),p should be understood as:
式(23-26)
Thus we can easily verify that the wave Iength in the horizontal direction of the disturbances are the same as that of the bottom configuration, and when k>ks, that is, when the bottom deformation is of very small scale, the amplitude of v1 is very small compared with those of ul and \omega\1, and the disturbances may
be considered to be confined to the x-z plane. When [Fig.107_038_08], the amplitude of v1 is still small compared with that of u1, and the variation of velocities in the vertical direction is periodic if the condition, [Fig.107_038_09],is satisfied.When [Fig.107_038_10], that is, when bottom deformation is large, the amplitude of v1 is large compared with that of \omega\1 and the disturbances may be corlsidered to pe almost * Strictly speaking, small deformation of the free surface is expected, as in the case of unusual internal waves. This effect, however, will be very small and may be neglected in this discussion.
horizontal, and when k < k_{\beta\}, which is the case of a very large scale deformation, \omega\1 << v1 and the disturbances are almost horizontal so far as the depth of water, H,is not so great.
These solutions coincide with those obtained by Queney (1947) if the depth H tends to infinity, and k_\beta\ tends to zero.
In general, when the condition pH << 1 is fulfilled,(that is, [Fig.107_039_01] )the disturbances may be simply written as:
式(27-30)
These solutions indicate that u1, and v1 may be considered as uniform in the vertical direction if the deformation of large scale is concerned. The above solutions can also be derived from (10) assuming that the water is homogeneous, u1and v1 are independent of z, and w1 is a linear function of z, satisfying the boundary conditions (17) and (18). In (30),the condition [Fig.107_039_03] is assumed, and the density distribution is just the same as suggested by Sverdrup. The effect of horizontal variation of density, however, is ineffective to the general pattern of the current, because we have assumed that the basic current is almost independent of the baroclinity of water.ln this respect we cannot discuss the influence of the density modification on the current pattern, which was the main part of Sverdrup’s theory. From (20), it is found that the stream lines deviate to the left near the summit of the bottom deformation in the northern hemisphere if k>ks and just the opposite if k < k_{\beta\}. This is illustrated in Fig.1.
- 幅:1840px
- 高さ:133px
- ファイルサイズ:9.8KB
- 幅:1146px
- 高さ:113px
- ファイルサイズ:12.2KB
- 幅:726px
- 高さ:100px
- ファイルサイズ:7.2KB
- 幅:600px
- 高さ:127px
- ファイルサイズ:5.9KB
- 幅:2086px
- 高さ:261px
- ファイルサイズ:17.9KB
- 幅:2033px
- 高さ:220px
- ファイルサイズ:15.5KB
- 幅:1999px
- 高さ:119px
- ファイルサイズ:10.4KB
- 幅:2146px
- 高さ:220px
- ファイルサイズ:17.6KB
- 幅:2313px
- 高さ:247px
- ファイルサイズ:22.4KB
- 幅:2160px
- 高さ:213px
- ファイルサイズ:18.4KB
- 幅:2546px
- 高さ:207px
- ファイルサイズ:24.8KB
- 幅:373px
- 高さ:227px
- ファイルサイズ:6KB
- 幅:2540px
- 高さ:773px
- ファイルサイズ:65.5KB
- 幅:363px
- 高さ:103px
- ファイルサイズ:3.9KB
- 幅:281px
- 高さ:164px
- ファイルサイズ:3.8KB
- 幅:259px
- 高さ:113px
- ファイルサイズ:2.7KB
- 幅:3046px
- 高さ:226px
- ファイルサイズ:47.5KB
- 幅:2106px
- 高さ:1093px
- ファイルサイズ:67.3KB
- 幅:794px
- 高さ:186px
- ファイルサイズ:10.8KB
- 幅:1572px
- 高さ:879px
- ファイルサイズ:56.6KB
4.Perturbations caused by the parabolic deformation of the bottom
Let us now investigate the case when bottom deformation is such that
[Fig.107_040_01]
and
[Fig.107_040_02]
We may write h by the Fourier integral as follows:
式(31)
Then, from (16),(17) and (18),we have:
式(32)
Again, from (10) and (12),we have:
式(33)
Both of these integrals can be evaluated by the method of contour integration in the complex plane.
Here we make the following transformations
式(34)
and
式(35)
where n is a positive integer. We may write approximately under the conditions [Fig.107_040_08]
式(36)
where
[Fig.107_040_10]
and
[Fig.107_040_11]
K.Kajiura:On the Influence of Bottom Topography on Ocean Currents.
In general, the conditions [Fig.107_041_01] are satisfied, and all these quartities are real.
Furthermore, we rewrite the terms of sine and cosine in exponential forms such as
式(37-38)
Thus, we must evaluate these integrals separately in each region, taking a suitable path of integration. The path may be taken from -R to R, excluding poles, on the real axis and upper or Iower semi-circle of radius R, so as to make the integral on the semi-circle to vanish whe:n R→\infty\. If we take the principal values in Cauchy’s sense, and substitute the conditions that v’=0 and w’=0 at x=-\infty\, definite integral values can be obtained. (See Lamb, H.:Hydrodynamics.6th Ed. P.408)
The results are:
式(39-41)
and
式(42-44)
where n0 is the largest positive integer less than ksH/\pi\, abbreviations are used:
[Fig.107_042_02]
[Fig.107_043_01]
From these solutions, we can conclude that \alpha_{n}\ and \delta_{n}\, which exist only when ksH > \pi\, contribute to establish the stationary internal waves in the down-stream side of the deformation of the bottom, and no conspicuous disturbances occur in the up-stream side. And these internal waves prevail in the current pattern when the scale of deformation, a, is sufficiently smal1(ak_{\lambda\}<<1), on the other hand, they may be negligible compared with the horizontal perturbations when the scale of deformation is large. These features can be expected already from the discussion of perturbations for the sinusoidal bottom profile.
For aks < 1, the terms containing \gamma_{n}\ and \delta_{n}\ are negligibly small compared with those of \alpha_{n}\ and \beta_{n}\ an, and, at a sufficiently large distance downstream, the terms of \alpha_{n}\ for n is near no predominate, since the terms of \beta_{n}\ will die out.
The leading term whern0 =1 is
式(45)
and the stream lines are given by
式(46)
For ak_{\lambda\}〜1, the terms of \alpha_{n}\ and \beta_{n}\ are rather small compared with the terms of \gamma_{n}\ and \delta_{n}\ , so that in this case, the whole terms in the equations must be taken into account.
For ak_{\lambda\} << 1, we may reasonably neglect the terms containing \alpha_{n}\ ,\beta_{n}\ \gamma_{n}\ and \delta_{n}\ and the solutions reduce to:
式(47-48)
and
式(49)
These solutions can be derived from the original equations if the several assumptions are introduced such that the water is homogeneous and the variation of u1 and v1 in the vertical direction is negligible. In these solutions, we can see at once that in the northern hemisphere, the stream lines deviate to the right by the planetary effect and it is followed by Rossby waves. The deviation of stream line is illustrated qualitatively in Fig.2.
- 幅:1080px
- 高さ:120px
- ファイルサイズ:6.8KB
- 幅:1073px
- 高さ:120px
- ファイルサイズ:9.4KB
- 幅:2353px
- 高さ:253px
- ファイルサイズ:23KB
- 幅:2626px
- 高さ:260px
- ファイルサイズ:30.7KB
- 幅:2939px
- 高さ:240px
- ファイルサイズ:33.9KB
- 幅:2206px
- 高さ:386px
- ファイルサイズ:32.2KB
- 幅:2179px
- 高さ:373px
- ファイルサイズ:32.9KB
- 幅:540px
- 高さ:126px
- ファイルサイズ:6.6KB
- 幅:2446px
- 高さ:253px
- ファイルサイズ:27.2KB
- 幅:2520px
- 高さ:426px
- ファイルサイズ:40.5KB
- 幅:2560px
- 高さ:420px
- ファイルサイズ:41.5KB
- 幅:719px
- 高さ:120px
- ファイルサイズ:7.9KB
- 幅:2966px
- 高さ:927px
- ファイルサイズ:100.6KB
- 幅:3072px
- 高さ:2033px
- ファイルサイズ:211.7KB
- 幅:3052px
- 高さ:2419px
- ファイルサイズ:253.4KB
- 幅:1580px
- 高さ:1532px
- ファイルサイズ:95.6KB
- 幅:1486px
- 高さ:1246px
- ファイルサイズ:74.1KB
- 幅:2146px
- 高さ:246px
- ファイルサイズ:19.9KB
- 幅:2126px
- 高さ:206px
- ファイルサイズ:18.4KB
- 幅:2779px
- 高さ:353px
- ファイルサイズ:38.2KB
- 幅:2753px
- 高さ:273px
- ファイルサイズ:28.5KB
- 幅:1559px
- 高さ:1013px
- ファイルサイズ:52.1KB
5. The Effect of horizontal shear of a basic current.
When the scale to be considered is large, and the basic current is constant in a vertical, we can assume from the above discussion that the horizontal velocities of perturbations are uniform in the vertical direction and the vertical velocity is a linear function of z. Therefore, on the assumption that the water is homogeneous and [Fig.107_044_03] ,eqation (8) becomes:
式(50)
provided that h is independent of y.
In a region of strong currents, it may be justified that \beta\ is negligible compared with d^{2}U,/dy^{2}. On these assumptions, we will discuss a very simple case, such as U=U0 cosk1y and h=h0cosk2x. A particular solution of the equation is then:
式(51)
where H is the depth of water.
In case when k1 is several times greater than k2,(k1=mk2),k2^{2} may be neglected compared with k1^{2}, and (51) reduces to :
式(52)
From the above solution,we can see that it is very important to take the horizontal shear of a current into account, if it is comparable with Coriolis parameter.
The larger the ratio of the scale of the bottom deformation to that of currents is, the greater the effect of the shear is.
For example, in case of
[Fig.107_045_01]
becomes 3 and the order of magnitude of m^{2}U0k1 is almost comparable with that of\lambda\,(〜10^{-4} sec^{-1}).
The current pattern is shown qualitatively in Fig.3.
- 幅:507px
- 高さ:179px
- ファイルサイズ:8.6KB
- 幅:2306px
- 高さ:253px
- ファイルサイズ:27.5KB
- 幅:2113px
- 高さ:260px
- ファイルサイズ:23.2KB
- 幅:2153px
- 高さ:227px
- ファイルサイズ:20.9KB
- 幅:1386px
- 高さ:466px
- ファイルサイズ:41.6KB
- 幅:1739px
- 高さ:1066px
- ファイルサイズ:70.6KB
6. Conclusions.
We have discussed above that when we are investigating the effect of bottom topography on a uniform basic current, it is important to consider the relation of the scale of the bottom configuration to various parameters such as ks,k_{\lambda\} and k\beta\, each of which is the representative scale of the motion of the stratified water on the rotating earth. When the scale of bottom deformations are very small, such that k_{\lambda\}a << 1, the parameter ks, which is related to the vertical gradient of the density, plays a dominant role and the perturbation may be almost two dimensional in the vertical plane, and, moreover, if ksH >\pi\, the internal gravity waves occur downstream, the wave length of which is 2H^{0}/n in the vertical direction and 2\pi\/\alpha_{n}\ in the horizontal direction.
When the scale of bottom deformations is very large such that k\lambda\a >> 1, the parameter ks, which is related to the latitudinal change of Coriolis parameter, plays a dominant role and the conclusions are the same as those obtained on the assumption of homogeneous water, such that the stream-lines deviate to the right in the northern hemisphere, accompanied by the Rossby waves.
In the case of an intermediate scale of phe:nomena, the parameter k\lambda\, which is related to Coriolis parameter itself, must be taken into account and the perturbations are considered as the ccmbined effects of various factors.
On the other hand, it was Shown that, if the basic current is not uniform, the horizontal shear of velocity may sometimes play an important role in the discussion of the topographic effect upon the currents. Thus, the deflection of currents may be related to the horizontal gradient of velocity and Coriolis parameter.
It must be remembered, however, that the whole discussions are made under the assumption that the steady perturbations always exist, regardless of values of various parameters, so that the discussions of stability and instability of these flow patterns are entirely ommitted.
7.Acknowledgements.
The author is indebted to Mr. H. Kusunoki, of the Low Temperature Science Institute, Hokkaido Univ., who generously communicated him the papers of V. W. Stockman. He also wishes to thank Dr. M. Nakano, who kindly read the manuscript before printing.
References
Bolin, B.,1950: 0n the influence of the earth’s orography on the general character of the westerlies. Tellus,2,(3), pp.184-195.
Charney, J. and Eliassen, A.,1949:A numerical method for predicting the perturbations of the middle latitude westerlies. Tellus,1,(2), pp.38-54.
Colson, D.,1949:Air flow over a mountain barrier. Trans. Amer. Geophys. Union,30,pp.818-830.
Ekman, V, W.,1923:Uber Horizontalzirkulation bei winderzeugten Meeresstromungen.
Ark. f. Mat., Astr. och Fys., Bd.17, No.26.
Ekman, V. W.,1932:Studien zur Dynamik der Meeresstromungen. Beitr. zur Geophys., Bd.36, S.385-438.
Queney, P.,1947: Theory of Perturbations in.Stratified Currents with Applications to Air Flow over Mountain Barriers. Misc. Report No.23, University of Chicago.
Queney, P.,1948:The problem of air flow over mountains;a summary of theoretical studies・Bull. Amer. Met. Soc.,29,(1), PP.16-26.
Stockman, V. B.,1047: Effect of Bottom Topography on the Direction of Currents in the Sea. Priroda.11, pp.10-23.(Defence Research Board, Canada., T 57 R.) Stockman, V. B.,1948:Effect of bottom topography on the direction of the transport of water set up by the wind or the muss-field in a non-homogeneous ocean. Dok. Acad.
Nauk. SSSR,59,(5), pp.889-892.(DRB, Canada. T 56 R.) Sverdrup, H. U.,1941:The influence of bottom topography on ocean currents. Applied Mechanics, Th. von Karman Anniv. Volume.
THE AGRICULTURAL AND MECHANICAL COLLEGE OF TEXAS Department of Oceanography and Meteorology College Station, Texas A FORCED WAVE CAUSED BY ATOMOSPHERIC DISTURBANCES IN DEEP WATER Kinjiro Kajiura Technical report 133-1 Project 133 - Reference 56-26T
Research conducted for the United States Department of commerce, 、Weather Bureau, Contract Ilo・ cwb-8717, through the Texas A&M Research Foundation. Presentation of material in this report is not considered to constitute final publication.
October, 1956
LIST OF FIGURES
Figure 1 Schenatic representation of each tern of Q
Figure 2 Air-pressure and mind stress distribution
Figure 3 Wind stress curl
Figure 4 Surface elevation due to pressure, \zeta_{1}\
Figure 5a Surface elevation due to wind stress
Figure 5b Surface elevation due to mind stress in sections parallel
to the y-axis
Figure 6 Deformation of the sea surface due to a model storm
Figure 7a Distribution of the xcomponent of the total flow
Figure 7b Distribution of the y component of the total flow
Figure 8a Distribution of \zeta\ along atraight lines (ky=0. +-0.04)
parallel to the x-axis
Figure 8b Distribution of Qx and Qy along atraight lines
(ky=0. +-0.04) parallel to the x-axis
Figure 9 Surface elevation, \zeta_{1}\ ,due to air pressure,
[Fig.107_048_01] , along the x’ and y
axis (ak=1)
ABSTRACT
The two dimensional forced wave in deep homogeneous water generated by a moving storm is discussed. The inverse barometric defororation of the sea surface is a fairly good approximation to the actual condition near the center of the storm, but the flow of water is quite different from that expected from the analogy of the one dimensional progressive wave. The geostrophic flow related to the wind stress curl is predomirant, and a permanent deformation of the water surface behind the storm is created, which is balanced by the geostrophic flow.
The critical case when the velocity of the storm movement is the same as the velocity of free waves in water, is briefly discussed. In this case, the two dimensional model gives quite a different picture coupared with the one dimensional model. It is found that a resonance condition does not obtain as in the simple one-dimensional problem for the critical speed. It is noteworthy that the two-dimensionality and the presence of Coriolis force will allow the establishment of an equilibrium height of the moving forced surge, even in the absence of bottom friction.
Introduction*
In deep rvater, it is generally believed that the response of the sea surface to the atmospheric pressure is inversely barometric and that the effect of wind stress is small compared with that of the former a t least in a one dimensional case. For a large scale wind system, however, the surface deformation due to stress becomes significant even in deep water.
Very few studies have been made to date on storm surges of two dimensions, because of mathematical difficulties, though the results derived from a two dimensional model might be quite different from those derived from a one dimensional model.
In this paper, a particular case of a two dimensional forced gave in deep homogenous water of infinite horizontal extent generated by a traveing circular atmospheric disturbance is studied and it is found for usual dimensions and speed of movement of a storm, that the quasi-steady solution for the surface deformation can be approximated, as is expected, by the equilibrium elevation of the sea surface near the center of the storm, and that the total flow is predominantly governed by wind stress curl and maintains geostrophic balance with a part of the deformation of the sea surface.
The present study excludes phenonena such as free inertiagravity waves and the readjustment of the density stratification in water which wouId play an important role in initial and boundary value problems.
*This work was originated from the necessity to understand the nature of waves due to atmospheric disturbances in an open sea so that we can formulate suitable boundary conditions at the edge of a continental shelf in connection with the study of storm surges in shallow water.
Basic Equations
The hydrodynamic equations of motion utilized are,
式(1-2)
and the equation of continuity is,
式(3)
Where x and y are the horizontal co-ordinates, t is time ,Qx, and Qy are the components of volume transport,Q,* in the x and y directions respectively, \zeta\ is the elevation of the sea surface from the undisturbed level,\tau_{x}\ and \tau_{y}\ are the components of wind stress , \tau\, in the x and y directions respectively, \zeta\ is the equilibrium height of the sea surface due to the atmospheric pressure, f is the Coriolis parameter, d is the depth of water, g is the acceleration of gravity, and \rho\ is the density of water.
The above equations are valid under the following assumptions:
1. The water is homogeneous.
2. Field accelerations are negligible.
3. The bottom friction is negligible.
4. \zeta\/d and d/ \lambda\ are both very small compared with unity where \lambda\ is the water wave length. This is the so-called long wave approximation.
Furthermore, in the following discussion the variation of the Coriolis parameter with latitude is neglected, and the depth is assumed constant.
[Fig.107_051_03] where u is the velocity vector in water and Z is vertical axis positive upward.
- 幅:2373px
- 高さ:567px
- ファイルサイズ:47.2KB
- 幅:2339px
- 高さ:320px
- ファイルサイズ:21.3KB
- 幅:553px
- 高さ:300px
- ファイルサイズ:7.9KB
Derived Equations
From the equation (1), (2) and (3), we obtain*
式(4)
where curl [Fig.107_052_02] ,and for the total flow, we have
式(5)
and
式(6)
where \nabla^{2}\ is the Laplacian operator with respect to x and y.
Now we can split \zeta\ into two parts, \zeta\ = \zeta\1+\zeta\2 ,such that equation (4) consists of two equation as follows:
式(7)
and
式(8)
Here it is obvious that \zeta\1 represents the surface elevation due to the atmospheric pressure gradient modified by the effect of wind divergence, and \zeta\2 represents the elevation due to curl of wind stress.
Corresponding to the terms \zeta\1 and \zeta\2, we can split Q into two parts, Q = Q1+Q2 and from the equations (5) and (6), we have,
式(9-10)
* \tau_{x}\ and \tau_{y}\ stand for \tau’_{x}\/\rho\ \tau’_{y}\/\rho\ respectively.
式(11)
and
式(12)
where we assumed that the ocean is initially in the undisturbed state, and in the two equations (11) ,and (12) ,we used equation (8) for modification of the original form.
Equations (7) and (8) have the form:
式(13)
where
[Fig.107_053_04]
respectively.
This type of equation is sometimes called the Klein-Gordon equation [Morse and Feshbach,1953, p.138] which appears in the Quantum Iiechanics, and the fornal solution is given in terms of a Green’s function (see Appendix A), but the actual integration is very complicated except for a case of special type of the forcing function.
Crease [1956] discussed the solution for the special case of a one diaensional forcing function.
The equations from (9) to (12) are the Poisson type, if \zeta\1 and \zeta\2 are known from equation(13). Inspecting the right hand sides of these equations we can easily find that Q consists of two parts;a divergent term associated with the timely variation of the sea surface, and a rotational term which is in geostrophic balance with the part of the elevation.
- 幅:3039px
- 高さ:320px
- ファイルサイズ:35.4KB
- 幅:1626px
- 高さ:233px
- ファイルサイズ:18.7KB
- 幅:2726px
- 高さ:287px
- ファイルサイズ:30.4KB
- 幅:2713px
- 高さ:246px
- ファイルサイズ:29.2KB
- 幅:2746px
- 高さ:340px
- ファイルサイズ:33.2KB
- 幅:2740px
- 高さ:326px
- ファイルサイズ:32.5KB
- 幅:2172px
- 高さ:520px
- ファイルサイズ:39KB
- 幅:2613px
- 高さ:286px
- ファイルサイズ:30.4KB
- 幅:2553px
- 高さ:306px
- ファイルサイズ:31.2KB
- 幅:2347px
- 高さ:270px
- ファイルサイズ:20.8KB
- 幅:2380px
- 高さ:520px
- ファイルサイズ:38.4KB
The Equation for a Forced Wave
If the external force is assumed stationary with respect to the co-ordinate system moving with a constant velocity, V , in the positive x direction, in other words, if the external forcing function can be expressed by
式(14)
a forced wave equation is derived from the equation (13) by means of the transformation,
[Fig.107_054_02]
and the result is,
式(15)
where x’ = x-Vt .
This equation excludes free waves which may play an important role in initial and/or boundary value problems. However, It is sufficient to get a forced wave solution if we are interested in a stationary wave moving with the storm in an ocean of infinite extent.
A Forced Wave Solution
We proceed to find a forced wave solution under the following assumptions.
The depth of the ocean, d , is large enough that the condition V^{2}/\epsilon_{d}\ << 1 is satisfied (see Appendix B).
The atmospheric pressure distribution is radially symmetric with respect to the moving storm center.
The wind distribution is assumed to be determined by the gradient wind formula. The effect of the movement of the storm - isallobaric effect - is neglected. Therefore, the distribution of wind stress is radially symmetric and has no radial component,(div\tau\ =0), (\tau\=\tau_{\theta\}\).
Tnus equation (15) can be written,
式(16)
where r , \theta\ are the polar co-ordinates with the origin at the moving storm center and \tehta\=0 on the positive x’axis. Also k^{2} = f^{2}/■
and
[Fig.107_055_01]
corresponding to, [Fig.107_055_02] respectively.
The solutions of homogeneous equation for equation (16) are given by modified Bessel functions of the first and ■ kind, I0(kr) and K0(kr), and the solution for the inhomogeneous equation is obtained by a standard method,such that
式(17)
where limits of integration are selected to satisfy the conditions, \zeta\ = 0 at r = \infty\ and \zeta\ = finite at r = 0 .
Solution (17) can be transformed by partial integraton,into a more convenient form for the practical computation,such that
式(18)
and
式(19)
Therefore, the elevation ,\zeta\2, is given by,
式(19’)
where \zeta\2 is assumed zero at the infinite distance in the positive
x direction.
In the solution for \zeta\1,the last two terms involving Bessel functions are considered to be the correction terms to the equilibrium elevation, and the magnitude of these terms depend on the shape and the scale of the air-pressure pattern. If the dominant wave nunber of \zeta\ is much larger than the wave number k (reciprocal of the Rossby’s deformation radius), these terms become very small.
In the solution for \zeta\2,we can easily find that the elevation due to curl \tau\ is inversely proportional to the speed of the storm movement and the depth of water. Therefore, in shallow water the effect of curl \tau\ increases and may exceed the barometric effect, provided the condition V^{2}/gd << 1 is satisfied. Because w is radially symmetric. \zeta\2 is not radially symmetric with respect to the moving storm center. In general, the permanent deformation of the surface is produced behind the storm, unless the following condition is fulfiled;
[Fig.107_056_01]
In the limiting case of k → 0 , which corresponds to the condition that the wave number of \zeta\ is much larger than the wave number k , solutions (18) and (19) are reduced to,
式(20)
and
式(21)
The solutions (20) and (21) are good approximations to the solutions (18) and (19) in deep water as will be shown later in a numerical example [see Appendix C for a simple physical interpretation of eqs(20),■
The solutions of the equations from (9) to (12) are (see Appendix D),
式(22-24)
and
式(25)
where
[Fig.107_057_03]
and
[Fig.107_057_04]
The last term in each solution represents the component of nondivergent flow which contribute to the geostrophic balance viith the elevation of the surface and the remaining terms represent nongeostrophic flow which contribute to build up the deformation.These terms are schematically illustrated in Fig 1, where the dominant term is that in geostrophic balance with \zeta\2・Near the storm area,the total flow is approximated within a few percent of error by,
[Fig.107_057_05]
and
[Fig.107_057_06]
where the P2 term is in geostrophic balance with the elevation (\zeta\1 - \zeta\), and (near the center of the storm) the last term is about ten times larger than the remaining terms.
- 幅:2272px
- 高さ:160px
- ファイルサイズ:13KB
- 幅:713px
- 高さ:300px
- ファイルサイズ:11.2KB
- 幅:2546px
- 高さ:313px
- ファイルサイズ:28KB
- 幅:2273px
- 高さ:247px
- ファイルサイズ:21.2KB
- 幅:1793px
- 高さ:478px
- ファイルサイズ:33.8KB
- 幅:607px
- 高さ:360px
- ファイルサイズ:10.4KB
- 幅:2753px
- 高さ:593px
- ファイルサイズ:50.7KB
- 幅:2746px
- 高さ:593px
- ファイルサイズ:48.4KB
- 幅:2766px
- 高さ:580px
- ファイルサイズ:49.6KB
- 幅:2286px
- 高さ:400px
- ファイルサイズ:24.1KB
- 幅:633px
- 高さ:380px
- ファイルサイズ:8.9KB
- 幅:1766px
- 高さ:219px
- ファイルサイズ:9.9KB
- 幅:2026px
- 高さ:334px
- ファイルサイズ:18KB
- 幅:2546px
- 高さ:793px
- ファイルサイズ:63.4KB
- 幅:2513px
- 高さ:273px
- ファイルサイズ:25.5KB
- 幅:1459px
- 高さ:1226px
- ファイルサイズ:58.2KB
- 幅:363px
- 高さ:183px
- ファイルサイズ:4.6KB
- 幅:1786px
- 高さ:327px
- ファイルサイズ:19.4KB
- 幅:2233px
- 高さ:340px
- ファイルサイズ:22.8KB
A Model Storm
The pressure distribution of our model storm is given by,
式(26)
where P is the atmospheric pressure, the suffixes \infty\ and o refer to the pressure at infinity and at the center of the storm, and R is a constant distance from the center where maximum wind is found [Hydrometeorological Report No. 31, 1954]
The conditions that the pressure and its gradient at r = a is continuous, determines k’ such that,
[Fig.107_058_02]
The gradient wind v_{\theta\} is governed by the equation,
式(27)
where \rho\’ is the air density and isallobaric terms associated with the movement of a storm are neglected. The solution of the equation (27) gives,
式(28)
where [Fig.107_058_05]
The wind stress formula is,
式(29)
where K is the friction coefficient for wind stress.
We can calculate curlZ \tau\ from the equations (28) and (29).
Nurmerical values used for evaluating \zeta\ and Q are:
[Fig.107_059_01]
- 幅:3059px
- 高さ:566px
- ファイルサイズ:54KB
- 幅:959px
- 高さ:340px
- ファイルサイズ:13.1KB
- 幅:2053px
- 高さ:307px
- ファイルサイズ:18.8KB
- 幅:2200px
- 高さ:320px
- ファイルサイズ:23.1KB
- 幅:553px
- 高さ:206px
- ファイルサイズ:7KB
- 幅:1860px
- 高さ:280px
- ファイルサイズ:13.9KB
- 幅:2332px
- 高さ:1619px
- ファイルサイズ:119.4KB
Numerical Results
The variation of atmospheric pressure and stress with radial distance is shown in Fig 2, and the variation of curl \tau_{\theta\}\ with radial distance is shown in Fig 3. The value of curl \tau_{\theta\}\ changes sign at about kr = 0.08 and is discontinuous at r = a , but the absolute value is very small except near the center of the storm.
The surface elevation due to pressure [\zeta\1,given by equation (18)] is shown in Fig 4 together with the equilibrium elevation (\zeta\).
Inverse barometric elevation is a very good approximation within a few per cent of error near the center (4 per cent at the center). A very slight depression appears at a large distance from the center and approaches zero at infinity, which is not shown in the figure.
The deformatio of the surface due to wind stress is quite different from the effect of pressufe gradient as shown in Fig 5, and the order of deformation due to stress is about one-tenth of that due to air pressure near the center.This elevation, \zeta\2,is inversely proportional to the speed of a storm, V,and in the present example,\zeta\2 is about 7 per cent of \zeta\1 near the center.
The total deformation of the surface is shown in Fig 6 which shows an almost inverse barometric deformation near the center of the storm and an unsymmetric character of elevation in the drection of storm travel. On. the lee side of the storm the permanent depression exists of the order of about 13-14 per cent of \zeta\maxs.
The distribution of the total flow Q is given in Fig 7 in temsof Qx and Qy. The predominant term is the geostrophic flow balanced by the elevation \zeta\2 and so the height contours given in Fig 4 may be interpreted as stream lines of the total flow to a good degree of approximation.
About 10 per cent of the total flow is in geostrophic balance with the pressure gradient caused by the difference between \zeta\ and \zeta\1,and few per cent is associated with the divergence field due to the movement of the disturbance.
The very small value of the flow associated with the divergence field compared with the geostrophic flow is explained by the fact that the depth of water is very large and a relatively small [Fig.107_060_01] is enough to build the surface deformation.
Sections of \zeta\ and Q along lines parallel to the x-axis and through the maximun wind region are drawn in Fig 8. Due to the prevailing effect of the geostrophic balance between current and elevattion, the reIattions between \zeta\ and Qx are quite different from
that expected for one-dimensionaI progressive waves.
- 幅:373px
- 高さ:140px
- ファイルサイズ:6KB
Discussion of the Results
If we remove the assumption that V^{2}/gd << 1,we can see from equation (15) that the forced wave will deviate considerably from the equilibrium elevation due to the effect of the storm movement. The critical case of gd=V^{2} will be briefly discussed in Appendix E.
In this case, the Coriolis force and the two dinensionality of the waves are important factors in determining the elevation of the sarface.
If we take into account the deviation of the wind direction from the gradient wind due to the friction at the sea surface, it is easily seen from equation (4) that the air-pressure effect will be modified by the convergcnce effect of air flow, but this effect is rather small in deep water.
On the otaer hand, the absolute value of curl \tau\ will decrease and the corresronding surface elevation,\zeta\2, and the total flow, Q2 will be considerably less than those given in the above example.
If the absolute value of wind velocity is the same, the values of \zeta\2 and Q2 will decrease by sin^{2}\gamma\ per cent due to the deviation, \gamma\; of the wind direction from the isobar.
As a result of these modifications due to the surface friction of air flow, the deformation of water surface will approach more and more to symmetric pattern and the equilibrium elevation of water will become a very good approzimation in deep water.
The effect of the isallocaric wind in the will stregthen the wind in the right falf and weaken the wind in the left falf of the storm and correspondingly the field of flow in water will be modified to strengthen the flow in the right half and weaken the flow in the left half.
The assumption made in the basic equations that the field acceleration (non-linear terms) is negligible,is verified by examining the order of these terms compared with the other terms. From the Fig 7a,b, we can see,
[Fig.107_061_01]
Therefore,
[Fig.107_061_02]
because, in general, [Fig.107_061_03] for deep water. Thus, we can neglect non-linear terms compared with the Coriolis term, at least in deeo water.
In the present paper, we discussed only the forced wave traveling with the storm, and the density of the water is assumed constant.
These two assumptions are not fulfilled in the actual ocean, where the phenomena are not quasi-steady, nor the density of water uniform.
The stratification of water in the ocean will affect the formation of the geostrophic flow through the readjustment of density distribution, and, moreover, the free waves of surface and internal modes, which are neglected in this discussion, play an important role to disperse the energy transferred from a storm to a wider region*.
The rigorous discussion should be based on the solution of an initial and boundary value problem, taking both inertia-gravity waves and density variation into consideration. The study in this direction is now continued.
*According to Veronis [1956], the ratio of baroclinic to barotropic energy is dependent on the scale of the flow and, when the scale is smaller than the barotropic radius of deformation as in the case of this paper, more energy goes into the baroclinic mode.
And the time of duration of the wind stress is an important factor in the generation of non-geostrophic motions.
- 幅:1259px
- 高さ:227px
- ファイルサイズ:14.3KB
- 幅:1287px
- 高さ:353px
- ファイルサイズ:19.9KB
REFERENCES
Crease,J.: Propagation of long waves due to atmospheric disrurbances on a rotating sea, Proc. Roy. Soc.,A.233,556-569, 1956.
Horse, P. M.,and H. Feshbach:Methods of Theoretical Phisics,McGraw-Hill,New York,p.138,1953.
Proudman, J.:Dynamical Oceanography, John Wiley &Sons, Inc.New York,p.295,1953
U.S.Weather Bureau: Analysis and synthesis of hurricane wind patterns over Lake Okeechobee,Florida,Hydrometeorological Rep.No.31,1954
Veronis, G.:Partition of energy between geostrophic and non-geo-strophic oceanic motiions,Deep Sea Res.,3,(3),157-177 1956.
Appendix A
Green’s function in the two dimensional infinite given by the solution of the foflowing equation:
[Fig.107_064_01]
The solution is,
[Fig.107_064_02]
where [Fig.107_064_03]
This Green’s function shows that the disturbance travel faster than the free wave velocity [Fig.107_064_04]
The solution for equation (13) is given by,
[Fig.107_064_05]
The limit of integration must be determined to cover all x0,y0, t0 provided c\tau\ > P.
- 幅:2840px
- 高さ:373px
- ファイルサイズ:38KB
- 幅:2453px
- 高さ:526px
- ファイルサイズ:40.9KB
- 幅:2553px
- 高さ:187px
- ファイルサイズ:18.3KB
- 幅:433px
- 高さ:153px
- ファイルサイズ:4.3KB
- 幅:3006px
- 高さ:494px
- ファイルサイズ:43.7KB
Appendix B
For d=1500 fathoms, and V=40knots,the ratio, V^{2}/gd, is O.02.
If the velocity of the storm, V,is small, the assumption, V^{2}/gd << 1,is valid even in shallow water. For example, say, for V= 10knots, V^{2}/gd=0.02, we have, d = 60 fathoms.
See also Appendix E.
Appendix C
The solutions (20) and (21) can be understood by the following simple interpretation.
In deep water the free wave velocity is much larger than the velocity of the storm Movement, the whole sequences may be considered as quasi-steady and quasi-geostrophic. Therefore, the atmospheric pressure gradient is balanced by the deformation of the sea surface \zeta\1 :
式(20)
Wind stress accelerates a water body such that,
[Fig.107_065_02]
andthe Coriolis force due to the flow Q2 wil be balanced by the pressure gradient due to the surface deformation,\zeta\2.
Thus, we have,
[Fig.107_065_03]
and from these equations together with the equltion of continuity, we have
[Fig.107_065_04]
and the solution is,
式(21)
The deviation from the geostrophic balance is included in the solutions (18) and (19) but the numerical example shows that, in deep water, the correction is small.
- 幅:1792px
- 高さ:260px
- ファイルサイズ:11.5KB
- 幅:447px
- 高さ:260px
- ファイルサイズ:6.5KB
- 幅:1773px
- 高さ:313px
- ファイルサイズ:21.4KB
- 幅:866px
- 高さ:233px
- ファイルサイズ:10KB
- 幅:2099px
- 高さ:326px
- ファイルサイズ:18.8KB
Appendix D
By virtue of the radially symmetric characterof \zeta\1,the right hand side of equation (9) can be written in the polar co-ordinate as follows:
[Fig.107_066_01]
Corresponding to tale above expression, we put
[Fig.107_066_02]
Then equation (9) can be split into three equations
式(1)’
式(2)’
and
式(3)’
The solution of equation (1)’ is given by,
[Fig.107_066_06]
The solution of equation (2)’ can be constructed from the elementary solutions of the homogeneous equation, r^{2} and 1/r^{2}. And the solution of equation (3)’ can be constructed from the elementary solutions of the homogeneous equation,r and 1/r.
If we take into consideration that P1 and P2 must be finite at r = 0 and r = \infty\ , we can obtain the required solutions,
[Fig.107_067_01]
and
[Fig.107_067_02]
The other equations can be solved similarly, if we take into consideration that [Fig.107_067_03] is radially symmetric in equations (11) and (12).
- 幅:2933px
- 高さ:613px
- ファイルサイズ:66.1KB
- 幅:1993px
- 高さ:294px
- ファイルサイズ:21.2KB
- 幅:2320px
- 高さ:313px
- ファイルサイズ:26.5KB
- 幅:2680px
- 高さ:353px
- ファイルサイズ:33.5KB
- 幅:2653px
- 高さ:360px
- ファイルサイズ:29.8KB
- 幅:593px
- 高さ:207px
- ファイルサイズ:5.8KB
- 幅:1326px
- 高さ:393px
- ファイルサイズ:17.1KB
- 幅:1073px
- 高さ:333px
- ファイルサイズ:14.1KB
Appendix E
For a critical case of gd=V^{2},equation (5) becomes,
式(1)’’ [Fig.107_067_04]
The solution of the above equation under the conditions \zeta\ = 0 for y=+- \infty\ is given by,
式(2)’’ [Fig.107_067_05]
where
[Fig.107_067_06]
Now we study the barometric effect only. Then,
[Fig.107_067_07]
Therefore, if we assume,
[Fig.107_067_08]
we have the following solution
式(3)’’ [Fig.107_068_01]
The second term in the bracket is the correction term to the equilibrium elevation and its magnitude depends on the scale of the storm, a , relative to Roesby’s deformation radius, 1/K .
For aK = 1 , which corresponds to a=200km, [Fig.107_068_02] =V=40 knots, and f =10^{-4}/sec, we have at the center of the storm,
[Fig.107_068_03]
On the other hand if we consider the case when all the quantities are constant in the y-direction, which is not the one dimensional case in a strict sense because we retain the quantity, Qy, the equation becomes,
式(4)’’ [Fig.107_068_04]
and assuming [Fig.107_068_05],we have
式(5)’’ [Fig.107_068_06]
At the center of the storm with aK= 1 , we have
[Fig.107_068_07]
For the case of the one dimensional forced wave without the Coriolis force, [Fig.107_069_01] gives the resonance condition and the elevation becomes indefinitely great [Proudman, 1953, P.295].
From the above consideration, it is clear that the Coriolis force and the two dimensionality of the waves are important factors in determining the elevation of the surface, and the deviation of the surface deformation from the equilibrium height becomes more and more conspicuous as the ratio [Fig.107_069_02] increases.
The solution (3)" for aK=1 computed numerically along the x’ and y axes is shown in Fig 9. The deviation of the surface deformation from the equilibrium shape is shown clearly.
- 幅:1326px
- 高さ:393px
- ファイルサイズ:17.1KB
- 幅:1073px
- 高さ:333px
- ファイルサイズ:14.1KB
- 幅:324px
- 高さ:141px
- ファイルサイズ:4.1KB
- 幅:2239px
- 高さ:320px
- ファイルサイズ:22.3KB
- 幅:2800px
- 高さ:639px
- ファイルサイズ:51.3KB
- 幅:3026px
- 高さ:847px
- ファイルサイズ:75.2KB
- 幅:214px
- 高さ:140px
- ファイルサイズ:3.2KB
- 幅:2672px
- 高さ:473px
- ファイルサイズ:38.9KB
- 幅:1966px
- 高さ:333px
- ファイルサイズ:18.8KB
- 幅:920px
- 高さ:226px
- ファイルサイズ:10.3KB
- 幅:2280px
- 高さ:320px
- ファイルサイズ:24.3KB
- 幅:667px
- 高さ:240px
- ファイルサイズ:7.2KB
ACKNOWLEDGEMENTS
This study is a part of the research on storm surges which is being continued, under the supervision of Mr. R.O.Reid, to whom the author is indebted for many valuable suggestions. The author also wishes to express his appreciation to Mr.J.D. Cochrane and Dr. G. . Heumann for their advice concerning the form of the manuscript. He is grateful to Mr. T. Williams for drafting of figures and Mrs. H. Proctor for the preparation of the report.
LIST OF SYMBOLS(Main symbols only)
[Fig.107_070_01] [Fig.107_070_02] ,vertical component of stress curl
d depth of water
[Fig.107_070_03] [Fig.107_070_04] ,divergence of stress vector
f Coriolis parameter
[Fig.107_070_05] external forcing function, eq (13) or (16)
g acceleration of gravity
[Fig.107_070_06] modified Bessel function of the first kind (order 0 and 1 respectively)
k [Fig.107_070_07], reciprocal of Rossby’s deformation radius,
eq (16)
K stress coefficient, eq (29)
[Fig.107_070_08] modified Bessel function of the second kind (order 0 and 1 respectively)
P1,P2 coefficients in eqs (22) and (23)
Q volume transport vector
Q1,Q2 volume transport vector corresponding to \zeta\1
and \zeta\2 resrectively
Qx,Qy components of volume transport in the x and y directions, respectively
r,\theta\ polar co-ordinates with the origin at the center of the moving storm; \theta\ is measured from the
positive x’一axis counterclockwise
S1,S2 coefficients in eqs (24) and (25)
t time
V velocity of the storm movement
w [Fig.107_070_09], vertical velocityof the movement of the
water surface due to curlz \tau\
x,y Cartesian co-ordinates, in which the positive x is in the direction of the storms movement.
x x -Vt , x co-ordinate relative to the moving storm
z vertical co-ordinate positive upward
\zeta\ elevation of the sea surface from the undisturbed level
\zeta\ equilibrium elevation of the sea surface due to atmospheric pressure
\zeta\ 1 elevation of the sea surface due to barometric effect including the effect of the wind divergence
\zeta\ 2 elevation of the sea surface due to wind stress curl
\xi\ [Fig.107_071_01]
\rho\ density of water (\rho\’ density of air at sea level)
\tau\ wind stress vector
\tau\x,\tau\y components of wind stress in the x and y directions, respectiveiy
\tau_{\theta\}\ tangential component of wind stress in the polar co-ordinate
[Fig.107_071_02] [Fig.107_071_03] Laplacian operator in the horizontal plane
Equation (16) is written as
式(A)
and the solutions of the homogeneous equation are given by modified Bessel functions of order zero, I0 (kr) and K0 (kr). The function I0 (kr) is finite at kr = 0 and infinite at kr = \infty\, and the function K0 (kr) is infinite at kr = 0 and zero at kr = \infty\. Therefore, corresponding to the boundary conditions that \xi\ is zero at r = \infty\ and \zi\ is finite at r = 0, we can construct a Green’s function such that
[Fig.107_075_02]
and
[Fig.107_075_03]
where [Fig.107_075_04]
From the above Green’s function, we have,
[Fig.107_075_05]
Some formula are listed below:
[Fig.107_075_06]
for large value of z,
[Fig.107_075_07]
- 幅:400px
- 高さ:160px
- ファイルサイズ:4.2KB
- 幅:500px
- 高さ:253px
- ファイルサイズ:8.4KB
- 幅:293px
- 高さ:159px
- ファイルサイズ:3.1KB
- 幅:493px
- 高さ:266px
- ファイルサイズ:8.6KB
- 幅:800px
- 高さ:173px
- ファイルサイズ:7.4KB
- 幅:800px
- 高さ:153px
- ファイルサイズ:8KB
- 幅:266px
- 高さ:227px
- ファイルサイズ:4.9KB
- 幅:800px
- 高さ:167px
- ファイルサイズ:8.7KB
- 幅:233px
- 高さ:246px
- ファイルサイズ:4.4KB
- 幅:480px
- 高さ:240px
- ファイルサイズ:6.8KB
- 幅:187px
- 高さ:227px
- ファイルサイズ:2.3KB
- 幅:493px
- 高さ:267px
- ファイルサイズ:8.8KB
- 幅:3326px
- 高さ:4264px
- ファイルサイズ:461.4KB
- 幅:3199px
- 高さ:2332px
- ファイルサイズ:283KB
- 幅:3039px
- 高さ:2325px
- ファイルサイズ:217.7KB
- 幅:2959px
- 高さ:2286px
- ファイルサイズ:252.3KB
- 幅:2179px
- 高さ:2272px
- ファイルサイズ:220.3KB
- 幅:2140px
- 高さ:280px
- ファイルサイズ:18.7KB
- 幅:1473px
- 高さ:199px
- ファイルサイズ:11.8KB
- 幅:1520px
- 高さ:206px
- ファイルサイズ:11.8KB
- 幅:2180px
- 高さ:206px
- ファイルサイズ:18.3KB
- 幅:3126px
- 高さ:313px
- ファイルサイズ:37KB
- 幅:1360px
- 高さ:500px
- ファイルサイズ:23.9KB
- 幅:1246px
- 高さ:946px
- ファイルサイズ:52.5KB
- 幅:2726px
- 高さ:4631px
- ファイルサイズ:453.5KB
- 幅:2653px
- 高さ:4664px
- ファイルサイズ:410.9KB
- 幅:3372px
- 高さ:3359px
- ファイルサイズ:453.4KB
- 幅:3332px
- 高さ:3299px
- ファイルサイズ:398.5KB
- 幅:3246px
- 高さ:2312px
- ファイルサイズ:286.1KB
- 幅:3286px
- 高さ:2279px
- ファイルサイズ:298.1KB
- 幅:2886px
- 高さ:4671px
- ファイルサイズ:427KB
A FORCED WAVE CAUSED BY ATMOSPHERIC DISTURBANCES IN DEEP WATER Kinjiro Kajiura The Agricultural and Mechanical College of Texas Department of Oceanography and Meteorology Technical Report 133-1 - Reference 56-26T October,1956
ERRATA AND SUPPLEMENT
Page Line or Equation Number Change
iii 5 Figure 5a should read, 「Surface elevation due to wind stress parallel to the y-axis.」
iii 6 Figure 5b should read, 「Surface elevation due to wind stress"
iii 11 and 14 Insert comma to read, 「(ky = 0, +-0.04)」
3 4 Should read, 「[Fig.107_082_01] stand for [Fig.107_082_02] respectively.For the total flow, we have.…」
3 Footnote Omit footnote.
4 4 「where we assumed that the ocean is initialy in the undisturbed State, and the last two equations (11) ,…」
4 eq (13) [Fig.107_082_03]
6 2nd para. 「The solutions of the homogeneous equation…… of the first and third kind,…」
6 last line, 2nd para.; 「… at r = C. (See Appendix F) 」
8 last two equations; [Fig.107_082_04]
8 last line; 「…… remaining terms except on the line ky = O,where [Fig.107_082_05] 」
9 eq (26) [Fig.107_083_01]
9 7 「Hydrometeological Report No,31,1954.The formula (26b) is assumed for r>a instead of the formula (26a) in order to have the finitet total volume of water due to the equilibrium elevation.
[Fig.107_083_02]
The formula (26a) becomes 〜R/r for a large value of r and the above integral becomes infinite. However, this formula is a very good approximation to a pressure pattern within a hurricane. Therefore,the distance a may be chosen to bo equal to the approximate radius of a storm area.
10 Numerical values : [Fig.107_083_03]
11 4 「… of storm travel. Behind the storm, there remains a permanent depression of the order …」
11 4th and 5th paragraphs to be replaced as follows:
The very small value of fhe flow associated with the divergence field conpared with the geostrophic flow is due to the assumption of deep homogeneous water. For the same surface deformation, the geostrophic total flow is prportional to the depth but the total flow associated with divergence field is not directly related to the depth and we can expect the predominant effect of the geostrophic term. This situation is quite different from the case of shallow water, where the total flow associated with a divergence field may overcome the geostrophic flow and thus the divergence term becomes very important in determinin a flow pattern.
Sections of \zeta\ and Q along lines through the maximum \tau_{x}\ region and parallel to the x-axis are shown in Figure 8, together with the section along the x-axis. Due to the prevailng effect of th e geostrophic balance between total flow and elevation, the relation between \zeta\ and Qx along the line, ky =0.04, is .quite different from that expected for one-dimensional model in which the condition Qx = V\zeta\ must be satisfied due to the equation of continuity.
21 List of Symbols should be placed in front of report, following page iv
Figure 1: (5) ■ Term should be turned 90° to the left.
Figure 8b: [Fig.107_084_01] should read [Fig.107_084_02] ,and the value of [Fig.107_084_03] should be plotted instead of
[Fig.107_084_04] , as follows:
[Fig.107_084_05]
- 幅:1859px
- 高さ:219px
- ファイルサイズ:17KB
- 幅:587px
- 高さ:200px
- ファイルサイズ:7.1KB
- 幅:960px
- 高さ:313px
- ファイルサイズ:9.8KB
- 幅:487px
- 高さ:247px
- ファイルサイズ:4.9KB
- 幅:413px
- 高さ:200px
- ファイルサイズ:5.4KB
- 幅:2640px
- 高さ:519px
- ファイルサイズ:41.7KB
- 幅:1073px
- 高さ:260px
- ファイルサイズ:13KB
- 幅:1873px
- 高さ:739px
- ファイルサイズ:46.2KB
- 幅:934px
- 高さ:294px
- ファイルサイズ:12.1KB
- 幅:880px
- 高さ:307px
- ファイルサイズ:11.8KB
- 幅:627px
- 高さ:266px
- ファイルサイズ:8.5KB
- 幅:800px
- 高さ:280px
- ファイルサイズ:10.3KB
- 幅:1980px
- 高さ:553px
- ファイルサイズ:35.5KB
EFFECT OF CORIOLIS FORCE ON EDGE WAVES (II) SPECIFIC EXAMPLES OF FREE AND FORCED WAVES(1) BY KINJIRO KAJIURA Department of Oceanography and Meteorology, Agricultural and Mechanical College of Texas
ABSTRACT
Modification of edge waves due to Coriolis force is examined in terms of two specific problems:dispersion of an initial deformation of the water surface ; the forced wave due to a moving atmospheric pressure disturbance. In both cases the scale of the disturbance is an important factor, and the larger the scale the larger the effect of Coriolis force. In the northern hemisphere, the initial deformation of the water surface is split into two free edge waves in which the one moving to the left (facing the coast from the sea) has a larger amplitude than that moving to the right. The contribution of Coriolis force to the edge wave amplitude for an initial deformation having a horizontal scale of several hundred kilometers is about 20% at the center in the initial stage. In the case of forced waves, movement of the atmospheric pressure system to the left is more favorable for exciting resurgent edge waves thanmovement to the right. The contribution of Coriolis force is about 10 to 20% in the forced period and 20 to 40% in the amplitude for an atmospheric disturbance of several hundred kilometers in diameter which moves with a speed of about 20 to 30 knots.
INTRODUCTION
Resurgent water level disturbances caused by hurricanes traveling parallel to a coastline were analyzed from the standpoint of forced edge waves by Munk, et al.(1956),who obtained periods of resurgences in close accord with some actual observations. However, the range of conditions available was not sufficient to fully establish the veracity of the theory. Later, Greenspan (1956) extended the theory to a more realistic situation by treating the transient problem with a Gaussian distribution for the pressure deficit in the atmosphere. One of the interesting features of Greenspan’s analysis is that, for large scale atmospheric pressure disturbances of size comparable to a hurricane, only the fundamental edge wave mode is excited while all other modes are negligibly small if realistic values for bottom slope and storm propagational speed are taken. He concluded that hurricanes are usually too big to produce the maximum surge height which is possible for a given pressure deviation.
The theory of edge waves so far has been applied to specific problems withoutCoriolis force, and this approximation is quite proper if the scale in space and time is reasonably small. However, the effect of the earth’s rotation should be taken into account if the period or the horizontal scale of the external force respectively are comparable to or larger than either 1/f or the deformation radius C/f (where C is a representative wave velocity or storm speed and f is the Coriolis parameter).
Reid (1958) has discussed the modification of free edge wave modes due to Coriolis force and has presented some general considerations in regard to forced edge waves. In the present paper, two specific problems are investigated in detail on the basis of the fundamental edge wave mode. First, the dispersion of an initial deformation of the water surface is studied. The propagation of free edge waves to the right and to the left (facing the coast from the sea) becomes unsymmetrical due to the presence of Coriolis force, and a greater fraction of the total energy is propagated to the left than to the right in the northern hemisphere. Next, the influence of Coriolis force on forced waves due to a moving atmospheric pressure disturbance is discussed, making use of a model similar to the one utilized by Munk, et al.(1956). In this case, movement of the pressure system to the left is more favorable than movement to the right for the development of resurgent waves so long as the velocity of the pressure system is less than the maximum group velocity of the edge waves. The contribution of Coriolis force amounts to about 10 to 20% for the period and 20 to 40% for the amplitude in the resurgent edge waves. The effect of wind stress is not considered in the present paper.
(1)Contribution from Department of Oceanography and Meteorology, A&M Colege of Texas, Oeeanography and Meteorology Series No.121.
Reptint from SEARS FOUND. JOURN. MAR. RES.
BASIC EQUATIONS
Assume that the coastline is straight and that the depth of water is linearly proportional to the distance offshore, with zero depth at the coast. Take the horizontal right-hand coordinates x and y so that the x-axis is directed offshore with the origin at the coast. Using the equation of continuity and the linearized equations of motion without friction [Reid,1958 :eqs.(1) to (3)], the equation for the surface elevation \eta\ (relative to the undisturbed level) due to atmospheric pressure disturbances can be shown to be
式(1)
where t is time, f the Coriolis parameter, g the acceleration due to gravity,\eta\ the equivalent height of water due to the atmospheric pressure deficit (relative to normal pressure), and s the bottom slope in the xdirection. The rotation \nabra^{2}\ is of conventional use with respect to xand y.
Auseful forced wave equation can be derived from (1), if the atmospheric disturbance moves with constant velocity V, parallel to the coastline, without change of pattern. Putting \eta\ = \eta\(x,\xi\) where \xi\ = y-Vt in (1) and assuming that all disturbances in the water are zero for t → \infty\ in the finite region, we find
式(2)
This relation is used in the second problem treated in this paper.
- 幅:2693px
- 高さ:486px
- ファイルサイズ:49.6KB
- 幅:2680px
- 高さ:487px
- ファイルサイズ:47.7KB
DISPERSION OF AN INITIAL DEFORMATION OF THE WATER SURFACE
Consider the situation in which the water body is initially at rest with a given surface deformation. For mathematical simplicity, we confine our attention to only the fundamental mode of edge waves by assuming a suitable deforlnation initially. specifically, the initial conditions are as follows:
式(3)
where Qx and Qy are the components of volume transport of water in the x and y directions respectively. The scale of the initial deformation is represented by a and corresponds to the distance, on the x or y axis, from the origin to the position where the height of the water surface is half the maximum height (see Fig.1). The unknowns \eta\ , Qx, and Qy for t > 0 can be expressed by the superposition of edge waves of the fundamental mode, in view of the form of the Fourier transform of \eta\ (with respect to y) at t = 0.
The normal mode edge waves may be found from (1), but they are not discussed here (Reid,1958). Edge waves of the fundamental mode for any longshore wave length (2\pi\/k) are given by
式(4a)
Figure 1. Graphs of the initial distribution of water level; upper graph is the relative water leveI along shore; lower graph is the relative water level offshore from the center of the initial disturbance.
式(4b)
式(4c)
where
式(5a)
式(5b)
式(5c)
Here Hi is the fundamental normal mode elevation (of unit amplitude at shore), Ui and Vi are the corresponding components of volume transport in the x and y directions respectively, and subscript j=1,2, or 3 correspond to the different frequencies \omega\1, \omega\2, and \omega\3. Later in this section, j=3 is omitted because there is no contribution from this mode to the surface elevation. Note that j=1and 2 correspond to waves moving in the positive and negative y directions respectively and that the phase velocity of the former is larger than that of the latter for the same wave length, though the group velocity is the same in both directions.
The functions \eta\, Qx, and Qy for t > 0 are formally given by the real parts of the following integrals
式(6a)
式(6b)
and
式(6c)
where、F(k) is the Fourier transform of \eta\ at t=0 and x=0, namely
式(7)
The weight functions p1 and p2 can be determined from the initial conditions such that
式(8a)
and
式(8b)
These weight functions p1 and p2 show that more energy goes to the negative y direction than to the positive y direction and also that the longer the wave length the larger is the difference in wave heights between two waves propagating in opposite directions, as shown in Fig.
2. The functions p1 and p2 are a special case of the pjn functions derived by Reid (1958).
Substituting (7),(8a) and (8b) into (6), we can split the elevation into four progressive waves:
式(9)
where
式(10)
式(11)
Figure 2. Graphs of the weight functions p1 and p2 versus relative wave number gsk/f^{2}, for the fundamental edge waves.
and
式(12)
In (9),\eta\1* and \eta\2* are the counterparts of the fundamenta1(Stokes) edge waves without Coriolis force traveling to the positive and negative y directions respectively, although the speeds in the positive and negative directions are different, as is evident in (12). On the other hand,\eta\1(f) and \eta\2(f) represent waves essentially due to Coriolis force.
At the center of the initial deformation x=y=0, it follows from (10) and (11) that initially(t=0)
式(13)
and
式(14)
provided that a << 4gs/,f^{2}. The latter condition implies that the scale a must be much less than 4000 km at midlatitudes for s=10^{-3}. The ratio of the elevations \eta\1,2(f)/\eta\1,2(\delta\) is expressed approximately by fa/C0, where [Fig.107_090_05] is the ordinary edge wave velocity with the wave length \lambda\=8a. The ratio computed for several values of a are shown in Table 1, from which we can clearly visualize the effect of Coriolis force. In the case of the initial deformation with an effective radius of about 400 km (a=100 km), the contribution of Coriolis force is greater than 20% of the ordinary edge wave. If we compare the waves traveling in the positive y and in the negative y directions, the difference of initial heights of the two progressive waves may become 40% of the average height of two waves. However, the resultant deformation of the water surface does not appreciably deviate from the deformation for the case without Coriolis force so long as time is not great, because \eta\1(f) and \eta\2(f) cancel each other for initial stages.
TABlE 1.
Making use of the method of stationary phase, we can integrate (10) and (11) approximately for large values of t, such that
式(15a)
and
式(15b)
The above solutions are valid for [Fig.107_091_04].
For fIyI, there exists no stationary phase in the integral given by (11). Practically speaking, the second condition requires that the time must be very long (beyond a practical limit) if fIyI is not very small, because ft>10^{3}IyI/gst. Therefore, if the stationary phase is attained at all, the value of fIyI/gst is already very small, and the wave amplitude is almost negligible near the origin. However, formally looking into the argument of the cosine term in (15a) and (15b) we can find that modification of the dispersive character of edge waves due to Coriolis force is such that
local period : 式(16a)
wave length : 式(16b)
phase velocity : 式(16c)
group velocity : 式(16d)
where
式(16e)
式(16f)
式(16g)
and
式(16h)
where the subscript o indicates the quantity without Coriolis force and the upper sign and the lower sign in (16a, b, c) correspond to the subscripts 1 and 2 respectively. The above relationships are all consistent with the discussion based on the normal modes (Reid,1958) such that C1 > C2 and G1 = G2 for the same wave length. The ratio of the wave amplitudes at a common distance IyI to the right and to the left is given by
式(17)
as shown in Fig.3. The amplitudes for both waves decrease very rapidly with decreasing a due to the exponential factor involved in (15a) and (15b).
As a whole, the effect of Coriolis force increases with increasing values of \alpha\,(\alpha\ < 1). This implies that, for a fixed time, the local wave length increases with distance from the origin, and consequently the effect of Coriolis force increases. However, for IyI > gst/f the condition of stationary phase does not exist and consequently the local wave length is undefined. It is probable that beyond this point, to the left, the water level decays monotonically to zero as in the outer part of the initial disturbance, but to the right the elevation is essentially zero beyond IyI = gst/f.
Note that the condition fIyI/gst<1, which is required to establish the stationary phase, gives the maximum group velocity G1,2 = gs/f; this maximum value corresponds exactly to the maximum group velocity derived from the study of the normal modes [Reid,1958:eq.(42)].
Figure 3. Ratio of the wave amplitudes at a common distance IyI to the right and to the left at time t,due to an initial static mound of water.
- 幅:2006px
- 高さ:347px
- ファイルサイズ:23.8KB
- 幅:1826px
- 高さ:186px
- ファイルサイズ:10.8KB
- 幅:2013px
- 高さ:1406px
- ファイルサイズ:77.3KB
- 幅:1853px
- 高さ:260px
- ファイルサイズ:15.8KB
- 幅:1860px
- 高さ:213px
- ファイルサイズ:14.1KB
- 幅:1919px
- 高さ:266px
- ファイルサイズ:16.3KB
- 幅:1926px
- 高さ:260px
- ファイルサイズ:16.9KB
- 幅:1879px
- 高さ:186px
- ファイルサイズ:8.9KB
- 幅:2253px
- 高さ:240px
- ファイルサイズ:18.8KB
- 幅:2266px
- 高さ:233px
- ファイルサイズ:26.1KB
- 幅:2293px
- 高さ:260px
- ファイルサイズ:27.1KB
- 幅:1653px
- 高さ:180px
- ファイルサイズ:10.2KB
- 幅:1986px
- 高さ:219px
- ファイルサイズ:15.8KB
- 幅:1960px
- 高さ:226px
- ファイルサイズ:16.8KB
- 幅:2040px
- 高さ:200px
- ファイルサイズ:14.5KB
- 幅:2493px
- 高さ:253px
- ファイルサイズ:23.7KB
- 幅:2646px
- 高さ:380px
- ファイルサイズ:33.7KB
- 幅:2026px
- 高さ:1606px
- ファイルサイズ:84.2KB
- 幅:2173px
- 高さ:240px
- ファイルサイズ:21.6KB
- 幅:1640px
- 高さ:239px
- ファイルサイズ:10.9KB
- 幅:2046px
- 高さ:273px
- ファイルサイズ:24KB
- 幅:508px
- 高さ:209px
- ファイルサイズ:7.6KB
- 幅:2673px
- 高さ:380px
- ファイルサイズ:44.3KB
- 幅:2667px
- 高さ:547px
- ファイルサイズ:61.8KB
- 幅:2693px
- 高さ:520px
- ファイルサイズ:62.4KB
- 幅:1293px
- 高さ:127px
- ファイルサイズ:13.6KB
- 幅:1806px
- 高さ:227px
- ファイルサイズ:13KB
- 幅:1813px
- 高さ:227px
- ファイルサイズ:12.9KB
- 幅:1800px
- 高さ:266px
- ファイルサイズ:17KB
- 幅:1799px
- 高さ:213px
- ファイルサイズ:11.9KB
- 幅:1626px
- 高さ:173px
- ファイルサイズ:9.9KB
- 幅:1646px
- 高さ:246px
- ファイルサイズ:13.1KB
- 幅:1646px
- 高さ:226px
- ファイルサイズ:12.7KB
- 幅:1660px
- 高さ:253px
- ファイルサイズ:13.8KB
- 幅:1893px
- 高さ:260px
- ファイルサイズ:18.4KB
- 幅:1907px
- 高さ:1586px
- ファイルサイズ:73.9KB
FORCED WAVE DUE TO A TRAVElING PRESSURE DISTURBANCE
For atmospheric pressure disturbances,(2) can be solved generally by means of the Fourier transform technique. However, for mathematical simplicity, we assume an atmospheric disturbance of the form
式(18)
This is the same form as that assumed by Munk, et al.(1956), which is convenient since it implies that we can confine our attention to fundamental waves only. The wave characteristics derived from this model may be considered as fairly representative of the actual situation because, according to Greenspan (1956), the fundamental mode of the edge wave is the only mode effectively excited by a pressure disturbance of the scale comparable to a hurricane.
Since \eta\ and \partial\\eta\/\partial\t are assumed zero for t → -\infty\ (which corresponds to \xi\→\infty\ forV > 0, and \xi\ → -\infty\ forV < 0), the Fourier transform of (2) with the pressure disturbance (18) may be solved straightforward, and the solution for \eta\ is formally given by
式(19)
where7. < 0 for V > 0 and \tau\ > 0 for V < 0 and [Fig.107_094_02],
The other parameters are
式(20a)
式(20b)
式(20c)
The quantities ks and G represent respectively the wave number for the resurgence without Coriolis force and the maximum group velocity.
If we assume f=0 (G→\infty\), the solution for \eta\ reduces to the case discussed by Munk, et. al.(1956) and by Greenspan (1956).
Making use of the contour integral technique, we can express (19) for V > 0 in the form
式(21a)
式(21b)
where
式(22)
For V < 0, we have a similar expression, provided IVI < G:
式(23a)
式(23b)
If IVI > G, the oscillatory part of the above solution is absent.
Asymptotically, for large values of [Fig.107_094_11], \phy\ is proportional to the atmospheric pressure disturbance \eta\. In fact, for IVI → 0, [Fig.107_094_12] for all values of x and I\xi\I.Thus, the forced response in the water consists of a train of resurgent edge waves of speed V behind the storm together with a disturbance restricted to the immediate neighborhood of the pressure system for which the water level is the direct effect of the atmospheric pressure. This result is qualitatively similar to that for the case without Coriolis force. However, if -V/G > 1,no resurgent wave exists in the present case. In other words, if the storm moves faster in the negative y direction than the maximum group velocity gs/f (this is very unrealistic), the resurgent waves do not develop. The case V = -G is the resonant condition and the water level increases indefinitely [Fig.107_095_01]. For the movement of the storm in positive y direction, no such condition exists.
The resurgent edge waves behind the storm are given by
式(24)
Therefore, the period of edge waves is given by
式(25a)
or
式(25b)
where T*(=2\pi\ lVI/gs) is the period of forced edge waves in the absence of Coriolis force. The contribution of Coriolis force in the forced period is approximately IVI per cent of T*if V is measured in meters per second, s=10^{-3} and f=10^{-4} sec^{-1}. Resurgences following a pressure pattern which moves to the right have a shorter period than that of forced edge waves without Coriolis force. The opposite is true of resurgences associated with a pressure pattern which moves to the left. If the scale of the pressure system is very large, the contribution of Coriolis force to the height of resurgent waves becomes large compared with that which occurs in the case without Coriolis force. Putting \eta\1 and \eta\2 as the amplitude of resurgent edge waves moving to the right and to the left (V > 0 and V < 0) respectively, and putting \eta\* as the amplitude of edge waves in the case without Coriolis force, we have
式(26)
Note that the ratio is independent of the slope of the bottom, and the effect of Coriolis force is measured by the number fa/IVI, which is what we anticipated in the beginning. For reasonable values of V and a (V〜 10 to 15 m/sec, a〜25 to 50 Km), the contribution of Coriolis force in the wave amplitude along the coast seems to be about 20 to 40%.
A somewhat similar integration is given by Lamb (1932:412).
The ratio of amplitudes\eta\1,\eta\2 and \eta\* to the equilibrium height \eta\0 along the coast is shown in Table II. There exists an optimum value of a for which the ratio becomes maximum if V and s are fixed.
The amplitude of edge waves decreases exponentially in the offshore direction, and the rate of decrease is larger for a storm moving in the positive y direction than that for a storm moving in the negative y direction.
TABlE II. RATIOS OF THE VARIOUS AMPLITUDES OF RESURGENT EDGE WAVES
RELATIVE To THE MAXIMUM EQUILBRIUM HEIGHT AlONG THE COAST FOR f=10^{-4}
SEC^{-1} AND s=10^{-3}. THE RATIO \eta\*/\eta\0 IS SHOWN IN ITALICS. UPPER FIGURES IN EACH BLOCK REPRESENT \eta\1/\eta\0 (STORM MOVING TO THE RIGHT) ;THE LOWER FIGURES REPRESENT \eta\2/\eta\0 (STORM MOVING TO THE LEFT).
- 幅:2060px
- 高さ:267px
- ファイルサイズ:20.2KB
- 幅:2120px
- 高さ:260px
- ファイルサイズ:23.5KB
- 幅:887px
- 高さ:120px
- ファイルサイズ:8.8KB
- 幅:1720px
- 高さ:133px
- ファイルサイズ:10.5KB
- 幅:1680px
- 高さ:133px
- ファイルサイズ:9KB
- 幅:1673px
- 高さ:126px
- ファイルサイズ:8KB
- 幅:1806px
- 高さ:140px
- ファイルサイズ:11.8KB
- 幅:2286px
- 高さ:153px
- ファイルサイズ:20KB
- 幅:2687px
- 高さ:320px
- ファイルサイズ:36.6KB
- 幅:2306px
- 高さ:140px
- ファイルサイズ:18.9KB
- 幅:2173px
- 高さ:147px
- ファイルサイズ:12.7KB
- 幅:767px
- 高さ:100px
- ファイルサイズ:7KB
- 幅:2200px
- 高さ:173px
- ファイルサイズ:17.9KB
- 幅:1913px
- 高さ:240px
- ファイルサイズ:16.6KB
- 幅:1893px
- 高さ:160px
- ファイルサイズ:13.4KB
- 幅:1787px
- 高さ:253px
- ファイルサイズ:14.2KB
- 幅:2753px
- 高さ:986px
- ファイルサイズ:101KB
CONClUDING REMARKS
In general, Coriolis force in the edge wave problem has the effect of producing unsymmetrical disturbances moving to the right and to the left along the coast, with waves moving to the left (negative y direction) from an initial static mound containing the larger fraction of total energy. For pressure-induced disturbances, a storm moving to the left is more favorable in the excitation of resurgent edge waves behind the storm than movement to the right. This is true so long as the speed of the storm does not surpass the maximum group velocity of the edge wave (order of 100 m/sec). The model utilized in this paper is not adequate for predicting changes in water level due to actual meteorological disturbance, because (besides the rather arbitrary pressure pattern in the present example) the effect of wind stresses is completely neglected. A brief investigation concerning the effect of wind stress (to appear later) shows that the stress may produce edge waves of the higher modes and that the cumulative effect of wind stresses parallel to the coast would remain behind the storm; this manifests itself as a geostrophic current system parallel to the coast (if the storm travels a long distance without change of pattern).
ACKNOWlEDGMENTS
This work is based in part on research conducted for the Texas A.& M.Research Foundation through sponsorship of the U. S. Weather Bureau (Contract Cwb-9281).
REFERENCES
GREENSPAN, H. P.
1956. The generation of edge waves by moving pressure distributions. J. fluid Mech.,1:574-592.
LAMB, HORACE
1932. Hydrodynamics,6th Ed., Cambridge Univ. Press.738 pp.
MUNK, WALTER, FRANK SNODGRASS AND GEORGE CARRIER
1956. Edge waves on the continental shelf. Science,123: 127-132.
REID, R.0.
1958. Effect of Coriolis force on edge waves (I) Investigation of the normal modes. J.mar. Res.,16:109-144.
高潮について 東京大学地震研究所 梶浦欣二郎
1.緒 言
最近十年間足らずの間に,世界各国において高潮に対する関心が急激に高まったようにみえる.その理由は,例えば日本では,1959年9月26日の伊勢湾台風の際に,高潮が原因となって名古屋を中心とする沿岸各地に大被害がひきおこされたので,防災的な見地から高潮の実態をよく知る必要にせまられたことがあげられるが,(1) ヨーロッパやアメリカにおいても同様なことがいえるようである.即ち,1953年1月31日から2月1日にかけて,冬期の低気圧が原因となってヨーロッパ北海に起った高潮の際には,オランダ附近で偏差約3mに達し,その時刻がほぼ満潮時と一致したために,オランダの大堤防も各所で決壊し,1421年の大洪水以来かってみなかったといはれる大洪水がおこって,南西部の島々は海水下に没するという大被害を生じ,その復旧には多大の苦労を要した.この被害の経験を二度と繰り返さないために,オランダ政府はその年の2月21日にデルタ委員会を設置してオランダ南西部の防災と開発に関する方策の研究を行い,委員会は壮大なデルタプランをつくり上げた.1957年頃からこの計画は着々と実行に移されつつあり,全計画の遂行には25年の歳月が見込まれている.(2) この高潮以来,北海をめぐる諸国で高潮に関する研究が多数発表きれている.(3) 一方アメリカでは1954年8月31日にハリケーンがニューイングランド地方を襲い,沿岸各地に災害をもたらしたが,特にナラガンセット湾の最奥部プロビデンス附近では最高水位約4.6mを記録した.この年には,数個のハリケーンが米国東海岸附近を通過し,ハリケーンの予報及びそれに伴う高潮の防御ということについて関心が高まったようである.(4)そこで,1955年の末にアメリカ気象局を中心とする国立ハリケーン研究プロジェクトが誕生し,その仕事の一部として高潮研究が強力に押し進められることとなった.(5)
勿論以前から高潮に関する研究は方々で行なわれ,特に日本では世界的に有名な20世紀初頭の本多等による日本諸港湾の静振の研究(6) 以来,伝統的に海水振動の研究はよく行なわれ,1934年9月21日の室戸台風あるいは1950年9月3日のジェーン台風による大阪湾の高潮などにっいては多くの調査研究がある.(7) イギリスにおいても,海の長波理論と関連して,DOODSON,PROUDMAN其他によるすぐれた理論的研究があり,また北海をめぐる各国では種々な高潮予報方式が試みられた.特にオランダではSCHALKWIJKが予報技術についての研究をまとめて1947年に一応完成させた.(8) アメリカでは,昔から東海岸やメキシコ湾沿岸がしばしば高潮に見舞われているが,目立った研究はなされていなかった.
ここで高潮研究の歴史を述べる余裕はないが,一般に世界の各国とも,予報に関してはその地域的な特性に応じた研究がなされ,得られた実験式は直接には他の国の高潮予報に応用出来ない.その理由は,第一に,外力として働く気象条件が異なること,第二に,海底及び海岸の地形が場所によって異なることである.例えば,アメリカのメキシコ湾沿岸や北東岸には一般に広い大陸棚があり,ハリケーンによって起る陸棚上の高潮が重要な問題であって,(9) 湾内の高潮は,外海に面したところの高潮にそれだけ余分に附加されたものと考えられる.これに対して日本では,台風によって起こる東京湾,大阪湾,有明海等のような湾内の高潮が問題となり,(10) 陸棚が余り広くないために湾口における高潮はそれほど高くならない.一方オランダやドイツ等のように,北海の奥部に位置する国々では,北海というかなり広くて限られた海の内部で冬季の低気圧がひきおこす高潮を対象とし,(11) イギリスでは,北海の外部から伝播して来る高潮にも注目している.(12) このように各国それぞれ特徴のある高潮研究が進められて来たが,最近に至って電子計算機の利用による流体力学方程式の数値解析法が発達し,世界中どこにでも利用出来るような高潮計算方式が開発されて来た.この場合に,個々の場合について変るのは具体的な気圧及び風の場の与え方と,地形の与え方とであって,階差方程式の解き方自体については一般性があり,どの国で発展した方式でも利用出来る.(13)
・本年4月22日に日本海洋学会と共催で開かれたシンポジウムで講演した要旨である.
2.高潮の定義
ここで,高潮という言葉が,海洋における異常水位の現象のうちのどれを指しているかをはっきりさせておく必要がある.普通には,気象的擾乱によって起こされた,数時間から数日にわたる急激な水位の異常現象を高潮と呼ぶようであるが,研究上の方便としては,標準的な検潮儀で記録された実測潮位と,推算潮位(天文潮)との差に長周期水位変動についての補正をほどこしたものを偏差と呼び,このうちで気象的擾乱によるものを高潮とすることが多い.註(1)このように定義された高潮には,潮汐とか,地震による津波は入らないけれども,周期の長いうねりや静振は含まれる可能性がある.そこで見掛けの周期からみて,数十分以下の短周期振動は全部除外した方がよきそうである.註(2) 長周期の側についても同様であって数日以上の周期をもつ変動は高潮から除外してはどうであろうか.註(3) その他,例えば海流の短期変動に応ずる水位変動(16) などは,間接的には気象的擾乱が原因であっても,普通の意味での高潮とは言い難いし,数時間の周期をもつ湾や陸棚の海水自由振動は,その原因は気象的な擾乱であるかも知れないが,一度外力がなくなって自由振動の状態になってしまえば高潮とは呼びにくい.このようにして,高潮を厳密に規定することはむつかしいが,もう一つ,本質的に高潮の現象を他の水位変動ときり離して考えられない場合がある.それは,浅海に於ける潮汐と高潮との非線型結合の可能性であって,この点からいうと,厳密には水位の変動全体の中から,高潮の部分を取り出すこと自体が不可能となる.本文ではごく常識的な定義にしたがって,台風や低気圧によって起こった,数時間から数日にわたる割合に目立った異常水位の現象を高潮と呼んでおく.
註(1) これはstorm surgeと呼ばれ,storm tideというときには潮汐をも含めて異常水位全体を指すことがある.
註(2) 和達,広野(1954)は長周期のうねりを高潮に含ませているが,長いうねりの外不連緕線や気圧波に伴なう波をも除外する.これらの波は短周期だと沖合では長波の性質をもたない可能性がある.
註(3) 例えばGroves・G・などの取扱った数日から一週間程度の水面変動は高潮に入れない方がよかろう.
3.高潮の高さ
台風やハリケーンのような激しい気象擾乱によってどの程度の高潮が起こるかを知るために,第1図には日本及びアメリカにおいて観測された高潮の高さを低気圧の最低気圧に対して示してある.註(4) この図でわかることは,同程度の台風とハリケーンとを比較すると,高潮の最高はメキシコ湾沿岸にあり,次いで北米東岸,最後に日本という順になるようで,これは地形によるものであろうと思われる.即ち,アメリカ南部メキシコ湾沿岸及び東岸のニューイングランド沿岸には広い大陸棚が発達しており,そこにナラガンセット湾のような細長い湾が附随しているが,日本の場合には陸棚が狭く,湾の入口近くでの高潮の高さがそれほど高くないことが主な原因である.一般に偏差が3mを越す高潮は珍らしく(北海の1953年の高潮で約3m),4mを越すものは余程特殊な場所に限られるようである.例えば日本の高潮で偏差3mを越すものは,室戸台風による大阪の高潮(3.1m),伊勢湾台風による名古屋の高潮(3.4m)であり,アメリカでは,最近の目立ったところで,1938年,1944年,1955年のニューイングランドの高潮及び1956年のメキシコ湾岸ルイジァナの高潮(19)があげられる.註(5)
ある一地点における高潮の時間的変化は,低気圧の規模,進路,速さ等によって大きく左右され一概には言えないが,ほぼ数時間から半日位の間に目立った水位の昇降が起こり,その前後にはゆるやかな水位変化のあるのが普通で,時には静振のような振動が後に附随することもある.第2図には,海岸線に対して種々な相対進路をとるハリケーンがあった場合に期待される水位変動を,実例によって示してある.普通の低気圧によるものでは,数日にわたる水位変化がみられることも多いが,高潮の高さ自体はさほど大きくない.不連続線に関係した水位の異常変化は,高潮の中に入れるか入れないかには問題があるが,一般にかなり短周期である.
広い大陸棚に面したほぼ直線状の海岸に対し,ハリケーンが海側から進行して来た場合,海岸に沿った各地の最高水位の分布が低気圧中心からの最短距離に対してどのようになるかを第3図に示す.勿論個々の場合についての違いは大きいけれども,平均的にはハリケーン進路の右側約数十粁のところで最大となるようで,中心から左右共に200km以上も離れると,高潮はかなり小さい.
このような高潮の性質をみると、高潮の最高波高については観測点附近(数十粁以内)の風の状態が最も重要な原因と考えてよいようであり,今までに多く提出された実験式では,Hを高潮の波高,\sigma\ を観測点附近の最大風速,△pを気圧偏差としてH=a△p十6U^{2}の型のものを与え,α及びbを決定することが多い .(20) この式は,高潮を気圧降下による海水吸い上げ作用と,風による海水の海岸への吸き寄せとによって大体説明出来ると考えていることに相当し,静的な見方と呼べよう.註(6)
これに対して,動力学的な見方では,陸棚上あるいは湾内の海水は一つの振動系をなしているために,外力の見掛けの周期に応じて力学的な増幅作用が働き,共鳴する場合には湾奥の波高が静的な作用のみを考えた場合にくらべて遙かに大きくなり得ると考える.註(7)
これらの静的あるいは動力学的な諸効果を明確に表現するのは流体力学の方程式であるから,次にはその最も簡単化した式によって話を進める.
註(4) 図中ハリケーンの最低気圧は公式から推定したもの,台風の場合には観測最低気圧が用いられているので,日本の高潮の諸点は全体として僅かながら右にずらして考える必要がある.メキシコ湾沿岸の高潮の高さには潮位の補正が行われていないが,潮汐の全振幅が30〜60cmの程度であるから大して問題にならないと考えられる.用いた資料は宮崎 (17)及びHarris (18) による.
註(5) インドにおいてもペンガル湾をサイクローンが北上するとき高潮が起る筈であるが詳細はわからない.
註(6) 実際にはa,bの中に平均的な動力学的効果が含まれていて, aは必ずしも1ではないが,個々の低気圧についての動力学的効果は入っていない.
証(7) このような動力学的功果を含んだ高湖予報の実験式をつくる試みがWILSON (1959)(21)によってなされた.
- 幅:2033px
- 高さ:2972px
- ファイルサイズ:164.9KB
- 幅:3293px
- 高さ:4117px
- ファイルサイズ:400.1KB
- 幅:2199px
- 高さ:1506px
- ファイルサイズ:90.4KB
4.高潮の方程式
よく知られているように,長波に関する運動方程式と連続の式とは次のように書かれる.
式(1)
及び
式(2)
ここに,tは時間,水平面上に座標軸x,yをとり,\nabra\はx,yに関するベクトル微分オペレータ・kは鉛直上方に向う単位ベクトルとする.Vは深さに対して平均した水平速度ベクトル,\zeta\ は平水面上の海面上昇量,hは水深・gは重力加速度,fはコリオリ因数である. \zeta\* は大気圧の空間的平均値からの偏差を水面の上昇量として表わしたもの,\tau\は海面風が水面に及ぼす応力,\tau_{b}\ は海水運動が海底に及ぼす応力(密度で割った量を用いる)である.註(8) この式では海水運動によるエネルギーの消耗はすべて海底摩擦によることになるが,\tau_{b}\ の解釈を拡大して水平渦動粘性によるものを含めてもよい.
さて,(1)は完全な運動方程式において,種々な仮定をおくことによって導かれたものであるから,その適用限界についての注意が必要である.仮定は,1)海水運動は小さく,その変化はゆるやかであるために,非線型の項はすべて省略出来る.註(9) 2)海水運動の鉛直加速度は重力速度に比して無視出来る.以上の仮定のうち 2)は波長の長い高潮のような海水運動ではごく海岸の近く(実際には水深と同程度の離岸距離)を除いては満足されていると思われるが,1)の仮定は水深がある程度ないと成立せず,遠浅な海岸で波高が水深に比して無視出来ない場合には非線型項を含む式を用いるべきである.また,防波堤などの設置によって,その開口部附近だけに速い流れの生ずるときには明らかに非線型の効果があらわれるものと予想される.このことは又浅海における潮汐と高潮との非線型結合についてもいえる.今は(1)、(2)についてのみ論じ,非線型項の影響については触れない.
高潮の性質を調べるのに最も簡単なモデルは,海底摩擦も地球自転による転向力も省略し,現象が準静的に起こっていると考えた時で,(1)は
式(3) [Fig.107_103_01]
となり,高潮は,風の吹き寄せ効果と気圧低下による海水吸い上げ効果とで表わせる.問題を一次元に限り,x=0 から x=L までの陸棚を考えてその間では水深も風速も一定とし,x=Lから外は水深無限大とすると,(3)から
式(4) [Fig.107_103_02]
となる. (4)は水面の昇り高は深きに反比例し海域の広さ(風域より小さい場合)及び風の応力に比例することを示し,浅くて広い陸棚(又は湾)に海岸向きの風が吹けば高潮の高さが最も大きくなり易いことがわかる.風の応力は一般に風速の2乗に比例するから,高風速では僅かな風速の違いも高潮の高きの変化には大きく影響する.今△Uの風速変化によって変化する高潮の高きを△\zeta\ とすれば(\zeta\*を0として) △\zeta\’\zeta\ =2△U/U となり,10%の風速変化に対して20%の波高変化が起こることになる.
地球自転による転向力が,高潮に対してどれくらいの影響をもつかは,無次元量 \alpha\=fL/Cの大小によってきまる.ここでLは海水運動の起こっている海域の長さ,Cは長波の平均伝播速度 [Fig.107_103_03], fはコリオリ因数である.もし \alpha\ < 1 ならば,地球自転の影響は考えなくてもよい.例えば, L〜60km, h〜26m, f〜0.8×10^{-4}sec^{-1} とすれば, \alpha\〜0.30となり波高に対して地球自転の影響がややきくという程度である.一方静振の周期Tに対しては,△Tを周期の変化として,ごく近似的に△T/T〜2/\pi^{2}\ (fL/C)^{2} となり,上述の例では△T/T〜0.02となって静振の周期に対して地球自転はほとんどきかないと見なせる.(22)
次に動力学的な効果に関して,重要な無次元昼は,まず \beta\^{2}=V^{2}/c^{2} であり,ここにVは風域の移動速度(強制波の速度)をあらわす.よく知られているように,(23) 無限に長い水道中の強制波は(1-V^{2}/c^{2})^{-1} の割合で増幅され, V/c=1 即ち強制波の波速が自由波の波速に等しくなると,増幅は無限大となる(勿論実際には摩擦の影響でこうはならない).これに対して,二次元的に拡がった強制波を考えると,増幅度は波の中心で [Fig.107_104_01] となる.今,地球自転の影響を考慮に入れると,一次元,二次元の場合ともにv/c=1においても無限大の増幅度を与えない.これは,V/c=1に近くなると転向力の効果が著しくなるためである.(24)
以上は無限に広い海の場合であるが,陸棚上あるいは湾内の高潮について考えるときは,陸棚(湾)の固有振動周期Tと風域の移動による強制波の見かけの周期T*との関係が重要となる.今,L,L*をそれぞれ陸棚(湾)の長さ,及び風域の大きさとすると
[Fig.107_104_02]
という無次元量 \gamma\^{2} が1に近づくと,即ち,v/C〜L*/4Lで増福度が最大になることが期待される. \beta\^{2} と \gamma\^{2} という二つの無次元量のうち,どちらが高潮の場合に重要であるかはL*/Lの大小によってきまり,L*/Lの小さいときは \beta\^{2}が,L*/L ≧ 2では \gamma\^{2} の効果が卓越する.
第4図には,陸棚上をある大きさの風域が進行する一つのモデルに対して数値計算した結果(25)が示され,一般の増幅度の目安として,静的な上昇量に対して, \beta\^{2} 又は\gamma\^{2}が1になったときの増幅度が右端に附け加えてある.これによるとL*/Lが1以上では,最大増幅度は1.5〜2倍である.
地形の影響に関しては,現象が複雑で簡単には述べられないが,定性的にいうと水深変化によって長波の部分反射,屈折,回折等が起こり,海岸では波が陸上にあふれるためにそのエネルギーの一部を失い,残りは反射される.複雑な過程によって生じた波は干渉し合うから全体としての波形は場所場所で相当に異なることが予想される.その上,風の応力の効果が水深に反比例することと,摩擦力も水深に関係することを考えると,個々の観測点について高潮の時間的変化を予測することは非常にむつかしい.
現在の高潮研究は,いかにして数値計算による高潮予報を行なうかということに集中しており,まず基本的には(1)をどう取扱うかが考えられている.勿論実際計算に際して非線型の項を導入することはさして困難ではない.ここで問題は,1)基本の徴分方程式を如何なる階差方程式で代用し,時間,空間の格子間隔をどのようにとり,計算をどう進めるかという,数値計算技術の問題と,2)風及び気圧の場をどのように与え,註(10) 海面及び海底の応力に対してどのような式を用いるか,及び海岸と沖合との境界条件をどうするかという高潮に固有な問題とに大別される.これらの問題について今までにも研究はあるが,まだ完全なものとはいい難い.(26)
ここでは,外海に於ける境界条件を如何に与えるかという問題に関連して,広い大洋上の海水運動について考えてみる.
註(8) 風の応力\tau_{y}\ はある一定の高さの風速Uの函数として [Fig.107_103_04] と与えられるが,k^{2} は厳密には常数ではなく,風速及び大気下層の安定度に関係する. また海底の応力 \tau_{b}\ は, \tau_{b}\=k^{2}V/V の形に書かれることが多いが,水面に風の応力が働く場合は流速の鉛直分布が変化するので表面応力に関係した項那必要となる.詳細は論じないが,海底摩擦が水深及び表面応力の水平分布によって変化することを念頭におく必要がある.
註(9) (波高)/(水深) << 1及び (波高)・(波長)^{2} /(水深)^{3} << 1が要求される.
註(10) 基本式が線型であれば,気圧及び風の場について「重ね合せ」の原理が適用出未るわけであるが,非線型となるとそうは行かない.
- 幅:2279px
- 高さ:253px
- ファイルサイズ:19.5KB
- 幅:2273px
- 高さ:193px
- ファイルサイズ:13.1KB
- 幅:2013px
- 高さ:120px
- ファイルサイズ:9.1KB
- 幅:2046px
- 高さ:220px
- ファイルサイズ:11.7KB
- 幅:331px
- 高さ:110px
- ファイルサイズ:3.3KB
- 幅:486px
- 高さ:133px
- ファイルサイズ:4.9KB
- 幅:1160px
- 高さ:134px
- ファイルサイズ:10.9KB
- 幅:2033px
- 高さ:2485px
- ファイルサイズ:169.3KB
東京大学地震研究所 梶浦欣二郎
5.低気圧による大洋の海水擾乱
今簡単のために,(1)及び(2)を変形し, \zeta\=\zeta\1+\zeta\2と置き,t=0 で\zeta\1-\zeta\2=0とする.
もし水深が一定であれば, c^{2}-gh と置いて次の式が得られる.(\tau_{b}=0\)
式(5)
及び
式(6)
もし海が極めて深いと,c^{2}が大きくなり,外力に対する海水の応答は迅速に行なわれて,(5),(6)の左辺の第二項,第三項が第一項にくらべて省略出来,右辺では風の応力の効果が水深に逆比例するので,これも小さくなり,結局 \zeta\1〜\zeta\*,\zeta\2〜0となって水面変動は気圧の違いによる吸上げにのみ依存することになる.一般に風の応力の収束 (一▽・\tau_{s}\,) は気圧傾度の収束と同様な作用をなし,風の応力の渦度は,本質的に地球自転の転向力を媒介として水面変動にあずかる.このとき渦度は積分効果としてきくので,作用時間が長くなるほど,渦度の水面変動に対する寄与は大きい.今軸対称な風の場をとると,中心に吹き込む風は,中心の気圧降下による水の吸い上げ作用を幾分助けるように働き,風の回転成分は,海水を軸のまわりに循環させるように働く.この海水の循環運動に対して転向力がそれと直角方向に働くから,転向力に釣合うような水面傾斜がもたらされる.従って,転向力がなければ水面傾斜も起らないということになる.註(11)
この考え方を実際の海洋に適用するには,海洋が成層をなしていることを考慮しなければならない.今簡単のために二層の海を考え,上層は下層よりも遥かに厚さが薄く、又上下層の密度の差は極めて小きい (10^{-3} 程度)とすれば,(5),(6)式で,水深hを上層の厚さh1 でおきかえることによって,\zeta\1及び \zeta\2 を近似的に躍層自体の変動量とみなすことが出来る.
この場合には,h1 << hであるから,躍層の変動に対する風の応力の収束と渦度の効果は大きくなるが,海面の変動に引き直してみると,まだ圧力傾度によるものにくらべて小さいのが普通である.しかしながら,海水運動について考えると,もし風系の移動速度がそれほど早くなければある一地点における風の作用時間が長くなるので,上層における循環流の流速は相当の大きさになり得る.即ち,大洋の上層では,圧値から傾度や風の応力の収束による海水運動にくらべて,風の応力の渦度による海水の循環流が卓越することが判る.
こう考えて来ると,大陸棚の先端部で,海面の上昇量を \zeta\1 で近似するだけでは,境界条件として不十分なことは明らかで,何等かの方法で流れの状態をも考える必要がある.一般に,成層した海における順圧と傾圧との運動の分離可能なのは,水深と,層の厚さが一定の場合に限られ,陸棚傾斜のように水深の変化するところでは,両モードの運動が相互作用をする.従って、大陸棚の端の境界条件は,これを式であらわすこと自体がむつかしい問題で,未だ解決していないといえる.
註(11) 今の議論は線型理論の範囲において成立ち,もし遠心力が大きいと,勿論転向力がなくとも循環運動に対応する水面傾斜は出来る.ここで,風の向きと海水の循環運動の向きが同じであることは,一見有名なエクマン理論に反するようにみえるが,実はこの循環流は傾斜流に相当し,エクマン理論によつて計算される吹送流は,丁度転向力に釣合うような水面傾斜をつくるために行なわれる海水運動に相当する.
- 幅:2806px
- 高さ:226px
- ファイルサイズ:21.3KB
- 幅:2832px
- 高さ:220px
- ファイルサイズ:21.6KB
- 幅:2026px
- 高さ:1566px
- ファイルサイズ:81.3KB
6.ハリケーン内の気圧と風との分布について
ハリケーン及び台風の気圧分布についてよく使用される実験式は次のようなものである.
高橋の式
藤田の式
Hydro.Met.U.S.
これらの三式は見掛け上異っているようであるが(p一p0)/(pn-p0)は r→\infty\ に対して等しく1-R/rに漸近する.今 p0,pn,R,という因数を同一とおいて比較すると第5図のようになり,実験式の違いは結局中心付近の気圧分布の違いということになる.アメリカでよく使われている式にはexponentialが入っているので,実測に合うようにp0,pn,Rを決めるにはややめのどうな点もあるが,Rは中心からほぼ最大風速地点までの距離を表わし,推定きれた中心気圧p0は割合に実際をあらわしているらしく,多くのハリケーンについて,これらの因数の値が計算されている.(27)
風速に関しては信頼すべきものは少ないが,1944年9月のハリケーンに対して,アメリカ気象局の水理気象課で,各所の観測値を綜合してつくったモデルは第6図(a),(b)のようなもので,風速は中心からの距離の函数として4時間ごとに与え,進行方向に対して直角に右側と左側とでは異った値をとらせる.その上,風の吹き込み角も,場所の函数として与えられているが,この方の精度は良いとは言い難い.今,以上の図をもととして,風の応力の収束と渦度とを計算してみると第7図(a),(b)のようになる.収束,渦度共に強さは似たようなもので,風域全体の広さに対して,かなり狭い範囲(中心から約50海里以内)に集中していることが判る.
- 幅:1939px
- 高さ:2974px
- ファイルサイズ:143.6KB
- 幅:1013px
- 高さ:113px
- ファイルサイズ:8.9KB
- 幅:1179px
- 高さ:140px
- ファイルサイズ:10.2KB
- 幅:839px
- 高さ:113px
- ファイルサイズ:7.5KB
- 幅:2086px
- 高さ:2852px
- ファイルサイズ:172.1KB
- 幅:2043px
- 高さ:1800px
- ファイルサイズ:113.3KB
- 幅:2133px
- 高さ:2066px
- ファイルサイズ:115.9KB
引用文献
1) 例えば,宮崎正衛,宇野木早苗,上野武夫:伊勢湾台風による高潮とその理論計算について.第7回海岸工学講演集,209-215,1960.
2) 例えば,「The Delta Plan」Information Dep’t of the Ministry of Transport and Waterstaat,15pp.,1958.
3) 例えば,H. Charnock and J. Crease:North Sea Surges. Science Progress,45,491-511,1957.
4) A.C. Redfield and A. R. Miller:Water levels accompanying Atlantic Coast Hurricanes. Met. Mono. Amer. Met. Soc.,2,1-23,1957.
5) H.F. Hawkins, Jr., L. G. Pardue and C. M. Reber:The National Hurricane Centor. Weatherwise,14,87-98,1961.
6) K.Honda, T. Terada, Y. Yoshida, D. Isitani:Secondary undulations of oceanic tides. Journ. College of Sci. Imp. Univ. Tokyo,24,113pp. and plates,1908.
7) 例えぱ,高潮一般について,K. Wadati and T. Hirono:Storm tides caused by Typhoons. Proc. UNESCO Symposium on Typhoons,31-48,1954.
8) W.K. Schalkwijk:A contribution to the study of storm surges on the Dutch Coast. Roy. Netherlands Met. Inst.,1258, pt.1,1947.
9) K.Kajiura:A theoretical and empirical study of storm induced water level anomalies. Tech. Rep. Texas A&M Research Foundation, Ref.59-23F,97pp.,1959.
10) M.Miyazaki:On storm surges which recently struck the Japanese Coast. Ocn. Mag.,9,209-225,1957.
11) (7)を見よ.
12) J.R.Rossiter:The.North Sea surge of 31 January and 1 Februarv,1953. Phil, Trans. Roy. Soc.London, A.246,(915.),371-400,1954.
13) W.Hansen:Theorie zur Errechnung der Wasserstandes und der Stromungen in Randmeeren nebst Anwendungen. Tellus,8,287-300,1956. G.W.Platzman:A numerieal computation of the surge of 26 June 1954 on Lake Michigan. Geophysics,6,407-438,1958.
G.Fisher:Ein numerisches Verfahren zur Errechnung von Windstau und Gezeiten in Randmeeren. Tellus,11,60-76,1959.
気象庁:伊勢湾高潮の総合調査,気象庁技術報告第4号,昭和35年7月,287頁.
気象庁:東京湾高潮の総合調査報告,気象庁及び東京都,昭和35年10月,247頁.
気象庁:大阪湾高潮の総合調査報告,気象庁技術報告第11号,昭和36年3月,235頁.
14) W.L.Donn and W.T. McGuinness:Air-coupled long waves in the ocean. Journ. Met.,17,515-521,1960.
15) G.W. Groves:Day to day variation of sea level. Met. Mono. Amer. Met. Soc.,2,32-45 1957.
16) D.Shoji On the variation of the daily mean sea levels along the Japanese Islands. J.Ocn. Soc. Japan,17,(3),1961.
17) (10)を見よ.
18)D.L. Harris:Meteorological aspects of storm surge generation. Proc. A. S. C. E., Hydr. Div., Paper 1859,25pp.,1958.
19) D.L.Harris:Hurricane Audrey Storm Tide. Nat. Hurricane Res. Project, Report No. 23,19pp.,1958.
20) 例えば,気象庁:潮位表昭和36年.
21) B.W. Wilson:The prediction of hurricane storm-tides in New York bay. Tech. Rep.Ref.59-20F, Texas A&M Research Foundation,120pp.,1959.
22) (9)を見よ.
23) J.Proudman:Dynamical Oceanography. John Wiley and Sons, Inc., New York,1953.
24) K.Kajiura Response of a boundless two-layer ocean to atmospheric disturbances:Dissertation, A&M College of Texas,139pp.,1958.
25) R. O. Reid:Approximate response of water level on a sloping shelf to a wind fetch which moves towards shore. Tech. Memo. No.83, Beach Erosion Board,47pp.,1956.
26) (13)を貝よ
27) (18)を見よ.
UGGI TSUNAMI COMMITEE ON THE PARTIAL REFLECTION OF WATER WAVES PASSING OVER A BOTTOM OF VARIABLE DEPTH Kinjiro Kajiura Earthquake Research Institute, The University of Tokyo, Tokyo, Japan
Abstract
Theories of partial reflection of water waves are reviewed. The cases of strong and weak reflections are treated and it is shown that the local reflection modulus, which is similar to the reflection coefficient for the case of a discontinuous change of the depth, plays an important role in the process of partial reflection of long waves, and that the contribution of the one-time reflection alone is enough to estimate the reflection coefficient for weak reflection. Some particular examples are worked out, and the variation of the reflection coefficient is investigated with regard to the ratio of the width of a transition zone to the wave-length of incident waves. It is found that if the ratio is less than about 0.05, the reflection coefficient may be approximated by the well-known formula for an abrupt transition, and if the ratio is larger than about 1, the reflection coefficient is quite small so that Green’s formula gives a very good estimate of the transmitted wave amplitude. As for the waves of arbitrary wave-lengths, the analytical work of Roseau is introduced and amplified to give some numerical results. The two-dimensional propagation of long waves is also discussed on the basis of Tatarski’s method on the statistical treatment of the wave propagation through an inhomogeneous medium.
Introduction
In order to understand the physical process which governs the behavior of tsunamis, various factors should be taken into account, In particular, on continental and insular shelves, and in bays and estuaries, complicated free oscillations of water, including certain kinds of progressive boundary waves, may be excited by waves advancing from the open ocean, and one of the most important aims of tsunami hydrodynamics should be to clarify the response characteristics of a water body in shallow water in relation to the incident wave train. In the present paper, however, we are mainly concerned with the propagation of waves in an open ocean and the partial reflection of waves due to bottom irregularities, so that no fixed lateral boundary is taken into consideration explicitly;the eigen oscillations and boundary waves generated near the coast, and the complicated interference of waves due to total reflection at the continental boundary are not within the scope of the present discussion,
Even if we neglect the effect of energy dissipation, and confine our attention to a linear approximation, the propagation of water waves over a bottom of variable depth is known to be complicated by such effects as partial reflection, refraction, and diffraction, in addition to the complication due to the dispersive character of water waves in the range of intermediate and short wave lengths. If the wave length \lambda\ is much larger and the wave amplitude a is much smaller than the depth d of water, provided \lambda^{2}\/d^{3} is small, the equation governing the propagation of water waves conforms to the so-called wave equation which is a linear differential equation of hyperbolic type, and many of the general properties of wave propagation are considered to be common to various types of waves encountered in nature irrespective of the special nature of the waves. Thus, the knowledge obtained in studies of electromagnetic waves, waves along electric transmission lines and filters, electron waves in wave mechanics, elastic waves and sound, etc.,can be utilized with advantage to the understanding of wave propagation in water, at least for waves of small amplitude and long wave length.
Although the general properties of a linear differential equation of hyperbolic type is very well known mathematically, general solutions become very complicated if the coefficients of the equation are not constant and it is not easy to discuss the details of the solution for each particular case. This applies in the case where the medium of wave propagation is not homogeneous, and various methods have been developed to handle this situation.Among them, numerical analysis, which has made great progress by making use of high speed computers, seems to be very promising for the discussion of wave propagation in an inhomogeneous medium. However, analytical methods of approximation are very useful from the standpoint of the physical understanding of the process of partial reflection, and they also help in formulating the program of a numerical analysis, and in interpreting the results in physical terms.
Reprint from lUGG monograph n 24
The one-dimensional propagation of waves in a continuous medium has been studied very extensively since the 19th century(see Rayleigh, Theory of Sound,1926), and the reflection and transmission coefficients of waves are given for the case of an abrupt change of the medium, and also for the cases of certain particular types of the variation of the medium. The problem of wave propagation in an inhomogeneous medium may be discussed for the following three cases separately. Firstly, the case in which the variation of a medium is confined within a very small fraction of one wave length of the wave in question. The case may be called the strong ref-ection. The other limiting case may be adequately called the weak reflection when the variation of a medium is gradual so that the change of a quantity in one wave length is small enough compared with the quantity itself. Between these two limiting cases lies the case when the variation of a medium occurs within a distance of half a wave length or so. Techniques employed to solve the problem are generally different for these three cases.
In the discussion of the one-dimensional propagation of waves, an approximation of the continuous change of a medium by the abrupt, discontinuous change of a medium is permissible at least for the case of strong reflection, and the formulas for the partial reflection and transmission of waves can be easily derived. For the case of weak reflection, various approximate methods such as W-K-B method can be applied and the well-known Green’s formula for the amplification of wave height in relation to the change of depth in water usually corresponds to the first approximation in the system of successive approximation. In the third case where the transition of the medium occurs within a distance of half a wave length or so, the problem should be solved without approximation in many instances and the solution is given for each particular case of the variation of a medium separately.
Here, it should be pointed out that water waves are somewhat different from the other kinds of waves described by the wave equation, because the long wave is only an approximation to the real motion in water when the vertical acceleration of a water particle is negligible compared with the hydrostatic pressure, and in the neighborhood where the abrupt change of a medium occurs, wave motions may not obey the hydrostatic law. This situation may be easily understood by considering the case of onedimensional wave propagation, when an infinitely thin barrier extends from the bottom to a certain depth of water. The partial reflection of long waves cannot adequately be described by the conditions of the continuities of the total flow and the water surface at the location of the barrier without the vertical acceleration being taken into account.
Thus, there is a possibility that the formula derived on the assumption of long waves may not be applicable to the case when the depth actually changes discontinuously. In reality, however, the depth change in the ocean is very gradual, that is, tie slope of the bottom does not exceed a few degrees in most cases of interest and the formula derived on the long wave approximation seems to be valid as long as the wave length is large enough.
If the wave length is not very large compared with the depth of water, the properties of wave propagation change significantly, and the conclusions obtained for long waves should be modified accordingly. For immediate and short ranges of wave lengths, the problem of wave propagation over a bottom of variable depth has been investigated by many able mathematicians and within the frame-work of a linear potential theory the problem of waves on a linearly-inclined bottom is said to have been solved completely (Stoker, Water Waves, Chap.5,1957). However, the partial reflection and transmission of waves over a varying depth is solved for very special cases only and it may be worthwile to give numerical result obtained so far.
As for the two-dimensional problem of wave propagation in a variable medium, the mathematical difficulty is so great that no adequate general theory yet exists as far as the author knows. For the case of weak reflection, however, it seems to be possible to discuss the two-dimensional propagation of long water waves over a bottom of variable depth from a fairly general point of view, by applying the method developed in the study of the propagation of electromagnetic and acoustic waves in the atmosphere, which is in the turbulent state, and in which the refractive index of the air fluctuates in random fashion. Tatarski (1961) has given a general treatment of the theory of wave scattering and of the theory of parameter fluctuations of short waves propagating in a turbulent atmosphere, and the results may be interpreted in terms of the statistical variations of the phase and amplitude in space of long water waves although the assumptions of weak reflection and randomness of the depth variation are severe restrictions to the application of the results to actual situations such as the propagation of tsunamis in the ocean.
Basic aproximate equations
Assume the fluid is ideal (no viscosity) and incompressible and the motion is irrotational. The disposition of the co-ordinate axes is taken in the usual manner, with the x, y plane the undisturbed water surface and the z-axis positive upward. The free surface elevation is given by z=r(x,y,t) and the bottom surface, by z=-h(x,y), We adopt the convention that letter subscripts attached to dependent variables denote differentiation with respect to the denoted independent variables, unless otherwise stated.
The equations governing the propagation of water waves of small amplitude are derived. in terms of the velocity potentia1\Phi\ in the following form:
式(1)
and
式(2)
式(3)
where g is the acceleration due to gravity and the subscript n denotes the differentiation normal to the bottom surface. The free surface elevation \eta\ can be determined from
式(4)
(2),and(4)are obtained by rejecting all but the linear terms in \eta\ and \Phi\ and their derivatives in the kinematic and dynamic free surface boundary conditions. The problem in this formulation belongs, from the mathematical point of view, to the classical boundary problems of potential theory.
On the other hand, the shallow water theory is an approximate theory which results from the assumption that the vertical component of the acceleration of water particles has a negligible effect on the pressure \rho\ so that the pressure p is approximately by the hydrostatic pressure. This kind of approximation is valid when it is assumed that the depth of the water is sufficiently small compared with some other significant length, such as, for example, the radius of curvature of the water surface, Thus the theory is sometimes called a long wave theory.
lf we assume the wave of small amplitude, the governing equations for long waves are
式(5)
式(6)
and
式(7)
where u and v are the velocity components in the x and y directions respectively. These equations are derived as the first order approximation of Euler’s original equations of motion and the equation of continuity by developing all quantities in terms of a small parameter \sigma\ (\sigma\ =d^{2}/l^{2} where d and l are respectively the representative lengths related to the depth acid the horizontal dimension).
The development of the higher order approximations are out of the present scope, but it is of some interest to mention that the solitary wave is obtained by the approximation of the second order. It is remarked that (5) to (7) are valid provided a/h << 1, h /\lambda\ << 1and a\lambda^{2}\ / h^{3} << 1 where a is the amplitude, \lambda\ the wave length, and h the depth of water. The importance of the last parameter a\lambda^{2}\ / h^{3} is pointed out by Ursell(1953) in the discussion of the long wave paradox in the theory of gravity waves.
The details of the approximate theory may be found in a text book (see Stoker, Water Waves, Pt.1,1957).
In terms of the integrated velocity Uh = U and vh = V,(5),(6) and (7) may be
written as follows:
式(8)
式(9)
and
式(10)
- 幅:2106px
- 高さ:167px
- ファイルサイズ:9KB
- 幅:2126px
- 高さ:153px
- ファイルサイズ:10.3KB
- 幅:2140px
- 高さ:153px
- ファイルサイズ:9KB
- 幅:2146px
- 高さ:133px
- ファイルサイズ:9.5KB
- 幅:2473px
- 高さ:140px
- ファイルサイズ:9.1KB
- 幅:2446px
- 高さ:140px
- ファイルサイズ:9KB
- 幅:2440px
- 高さ:160px
- ファイルサイズ:10.9KB
- 幅:2440px
- 高さ:133px
- ファイルサイズ:9.4KB
- 幅:2460px
- 高さ:133px
- ファイルサイズ:9.5KB
- 幅:2493px
- 高さ:147px
- ファイルサイズ:10.5KB
One-dimensional propagation of long waves
For the propagation of long waves in a canal of variable cross section, the equation of motion is
式(11) [Fig.107_114_01]
and the equation of continuity is reduced to
式(12) [Fig.107_114_02]
where b is the width of a canal, Q( =bU) is the volume flux, and the variations of \eta\ and U across the canal are assumed to be negligible. If we set b constant, all the subsequent treatments will be reduced to the ordinary discussion of the one-dimensional propagation of waves in which the effect of the varying width is not taken into account.
The wave equation derived from (11) and (12) is
式(13) [Fig.107_114_03]
or
式(14) [Fig.107_114_04]
If we assume \eta\ and Q to be periodic with respect to time (e^{iwt}),(13) and (14) may be written
式(15) [Fig.107_114_05]
or
式(16) [Fig.107_114_06]
In later discussions, the time factor is ommittee for simplicity.
In general, the one-dimensional wave equation can be written for the case of
periodic waves in the form:
式(17) [Fig.107_114_07]
This is the equation encountered in various problems related to waves, and it is very well investigated in the field of applied mathematics. The general properties of a solution change according to the sign of k^{2}(x), and when k^{2}(x)is a slowly changing function of x, various methods of approximation are applicable, among which the socalled W-K-B method is as follows:
Set [Fig.107_114_08] and substitute into (17). If we consider [Fig.107_114_09] as a correction term in the corresponding equation of Riccati type,
式(18) [Fig.107_114_10]
the first approximation of the original equation is found to be
式(19) [Fig.107_114_11]
The second approximation is givenby
式(20) [Fig.107_114_12]
This solution constitutes an actual W. K. B, solution, and corresponds to Green’s formula for the long wave propagation without reflection.
First of all, we treat the case of strong reflection when the change of the medium is abrupt. The conditions at the discontinuity of the medium are given by the continuities of the elevation \eta\, and the volume flux Q. In terms of Q or \eta\ alone, the conditions are
式(21) [Fig.107_114_13]
or
式(22) [Fig.107_114_14]
where the subscripts 1 and 2 denote,respectively, the quantities in two regions x>O and x=> 0 of the medium separated by the discontinuity at x=0 (Fig,1).
Since b and h are constant for each region, the solution for Q of (16), which gives progressive waves moving in the negative x direction for x < 0, is given by
式(23) [Fig.107_114_15]
and
式(24) [Fig.107_115_01]
where
[Fig.107_115_02]
The boundary conditions (21) determine the relation between the constants A,Band C.
The reflection coefficient may be defined by
式(25) [Fig.107_115_03]
This formula is very well known and forms the basis of the subsequent discussion
(Lamb, Hydrodynamics,§176,1932).
Fig.1
[Fig.107_115_04]
- Abrupt transition.
Now, to discuss the partial reflection of the wave in general, it is more convenient to formulate the problem in terms of the parameter related directly to the reflection of the wave. Studies in this direction have already been persued by Pierce (1943)who obtained a first-order differential equation for the input impedance of a non-uniform electric line, and by Walker and Wax (1946) who derived a similar equation for the voltage reflection coefficient of non-uniform lines. The application of the method to the long water wave problem was made by Ogawa and Yoshida (1959).
Take solutions of the form:
式(26) [Fig.107_115_05]
and
式(27) [Fig.107_115_06]
where k^{2} =\omega^{2}\/(ghb^{2}). The substitution of (26) and (27) into (11) and (12) yields the equation for the reflection coefficient R defined by
式(28) [Fig.107_115_07]
as follows:
式(29) [Fig.107_115_08]
The reflection coefficient R defined by (28) is for Q of the waves moving in the positive x direction.The reflection,coefficient R’ for \eta\ can be easily defined in similar manner,
式(30) [Fig.107_115_09]
and the equation for R’ becomes
式(31) [Fig.107_115_10]
As is pointed out by Schelkunoff (1951), no meaning can be attached to the reflection coefficient in general when the incident wave is in an inhomogeneous medium for the simple reason that we cannot separate the total wave into 「progressive」 components.
A formal expression such as
[Fig.107_116_01]
in the form of an apparent「progressive wave」function in a homogeneous medium does not insure that it represents a wave with「progressive」physical characteristics. Thus, we shall restrict the definition of the reflection coefficient given by (28) or (30) to conditions where the incident wave travels in a homogeneous medium and impinges on an inhomogeneous medium of width l at x=x1. From [Fig.107_116_02] to x=\infty\ we assume another homogeneous medium to recognize the transmitted wave. Under some conditions, we may be able to let l approach infinity.
In general,(29) and (31) are valid, even if the variation of the medium is not gradual, but the equation is non-linear and it is not so simple to solve the Riccati equation unless R^{2} is small in comparison with unity, or, in other words, the solution is easily obtained for weak reflection only.
If we neglect R^{2} compared with unity, the solution of (29) is given by
式(32) [Fig.107_116_03]
where x2 is the position when R is assumed to be zero. The physical meaning of this expression becomes clear if we consider that [Fig.107_116_04] is analogous to the reflection coefficient (25). Essentially, the above solution (32) shows that the one-time reflection of progressive waves contributes to the main part of the reflected waves as a whole for the case of weak reflection (Ogawa and Yoshida,1959).
Since the expressions (26) and (27) formally correspond to the first order approximation (19) of the wave equation (17), it is probable that the second order approximation (20) may yield a better result in the discussion of the wave reflection. Actually, it turns out to be the case. The details of the development are given by Bremmer(1949, 1951) for electromagnetic waves;he showed that the solution of the wave equation can be expressed by a series of terms, with a simple physical meaning, the first of which constitutes the W-K-B approximation and the successive terms are interpreted as a successively higher times reflection of the incident waves, The outline of the method is given below in terms of long water waves.
Instead of (26) and (27), we assume
式(33) [Fig.107_116_05]
and
式(34) [Fig.107_116_06]
The substitution of (33) and (34) into (11) and (12) yields,
式(35) [Fig.107_116_07]
and
式(36) [Fig.107_116_08]
Now set
式(37) [Fig.107_116_09]
and
式(38) [Fig.107_116_10]
and consider \beta\(x) to be a quantity of order unity.
Then, we may write
式(39) [Fig.107_116_11]
and
式(40) [Fig.107_116_12]
These equations clearly show that the wave is partially reflected. The waves travelling in the positive x direction and in the negative x direction change their amplitudes respectively according to the partial reflection of the waves moving in the opposite direction, Thus \epsilon\\beta\1 and \epsilon\\beta\2 may be considered as a kind of local reflection modulus per unit distance at a particular location. The absolute value of the local reflection modulus thus defined is consistent with the well known reflection coefficient (25). This view enable us to interprete (32) as a consequence of the one-time reflection of incident waves.
Now, if \epsilon\ is less than unity, the solutions of (39) and (40) may be expressed by
式(41) [Fig.107_117_01]
and
式(42) [Fig.107_117_02]
and these series converge rapidly if \epsilon\ is small, that is to say, for the slowly changing depth.
Here we may put zero order terms to be constant so that
[Fig.107_117_03]
This is the case of no reflection and corresponds to the solution (20) for W-K-B approximation. The first order correction terms A(1) and B(1) are to be expected as representing secondary waves originating from reflection losses of the zero-order wave B(0) and A(0), respectively. In a similar manner,corresponding consideration applies to the connection between the (n-1)th term and the nth term because we may write
式(43) [Fig.107_117_04]
and
式(44) [Fig.107_117_05]
If we follow the same argument to the zero order wave A(0), B(0), it is seen
that the nth term in the above development represents the field contribution of a special wave that originates from \eta\ successive reflections of the zero order wave due to the inhomogeneity of the bottom. It can be shown from the mathematical point of view that the series solutions of the form (41) and (42) are actually the solutions of (39) and (40).
Here it may be noted that for weak reflection the deviation of the transmitted wave amplitude from that expected from Green’s formula is one order of magnitude smaller than the reflected wave, in other words, Green’s formula gives a very good estimate of the transmitted wave amplitude even if the partial refiection exists on a sloping bottom (Reid,1957). An elementary explanation may be given as follows:
Assume the incident wave amplitude A and the reflection coefficient \epsilon\. Then the reflected wave amplitude is of the order \epsilon\ A and the conservation of energy requires the transmitted wave amplitude x to be x^{2} A^{2}(1-\epsilon^{2}\). Therefore, we have x A(1- \epsilon^{2}\/2) for small \epsilon\,and the percentage variation of the transmitted wave amplitude is
[Fig.107_117_06]
This is one order higher than the reflection.
In connection with the above stated series solution in powers of \epsilon\,(41) and (42), we may have a wave equation explicity in cluding \epsilon\. If we set
[Fig.107_117_07]
where k^{2}=\omega^{2}\/(ghb^{2}) as before, the equations (15) and (16) may be transformed into
式(45) [Fig.107_117_08]
and
式(46) [Fig.107_117_09]
Assuming the local reflection modulus to be
式(47) [Fig.107_117_10]
(45) and (46) can be written as
式(48) [Fig.107_117_11]
and
式(49) [Fig.107_117_12]
It is to be noticed that the variable \xi\ is proportional to the number of wave lengths comprised between x=0 and x=x. In other words, the distance is measured with the aid of the local wave length as the unit. The quantity \epsilon\\beta\ may be interpreted as the reflection modulus of a transition zone per 1/(bk) and actually corresponds to the well known reflection coefficient (25) for the discontinuous change of the medium. These equations (48) and (49) are useful in understanding the physical process of the partial reflection because the solution can be expressed in terms of the reflection modulus, and this point is illustrated by an example later.
If we have discontinuous changes of the medium at some points, the boundary can be interpreted as the continuities of [Fig.107_118_01] or the continuities of [Fig.107_118_02].
The method of characteristics which is originally developed for the solution of non-linear waves such as shock waves (Courant and Friedrichs, Supersonic Flow and Shock Waves,1948) is quite useful to solve the linear wave problem by numerical means as is pointed out by Reid (1957).
Starting from the equations (11) and (12), the following set of equations is derived.
式(50) [Fig.107_118_03]
and
式(51) [Fig.107_118_04]
where
式(52) [Fig.107_118_05]
and
式(53) [Fig.107_118_06]
and
式(54) [Fig.107_118_07]
式(55) [Fig.107_118_08]
This set of equations (50) and (51) is quite similar to the set of equations (39) and (40), and Mx/M corresponds exactly to kx/2k ,showing the local reflection modulus per unit distance. The details of the discussion related to (50) and (51) are not given here.As shown by Reid (1957), the numerical computation of the wave propagation can be performed by means of a finite difference scheme and in this formulation the local reflection modulus will become
[Fig.107_118_09]
Instead of Mx/M wave length \lambda\ and the distance of ghe grid points Δx selected for numeerical computation becomes an important factor.This kind of modification factor,sin(2\pi\Δx/\lambda\)/(2\pi\Δx/\lambda\),is common to the finite difference scheme applied to the continuous system and the waves defined by the finite difference equations derived from the equations of a continuous system exhibit behavior somewhat analogous to the waves in a discrete system such as the electromagnetic waves transmitted through a lattice structure.
- 幅:2480px
- 高さ:133px
- ファイルサイズ:10.1KB
- 幅:2479px
- 高さ:167px
- ファイルサイズ:11.3KB
- 幅:2493px
- 高さ:153px
- ファイルサイズ:11.5KB
- 幅:2473px
- 高さ:160px
- ファイルサイズ:12.5KB
- 幅:2513px
- 高さ:206px
- ファイルサイズ:14.9KB
- 幅:2513px
- 高さ:207px
- ファイルサイズ:16.2KB
- 幅:2466px
- 高さ:147px
- ファイルサイズ:11.2KB
- 幅:553px
- 高さ:86px
- ファイルサイズ:4.6KB
- 幅:133px
- 高さ:100px
- ファイルサイズ:1.5KB
- 幅:2474px
- 高さ:146px
- ファイルサイズ:11.3KB
- 幅:2480px
- 高さ:180px
- ファイルサイズ:15.4KB
- 幅:2458px
- 高さ:237px
- ファイルサイズ:18.8KB
- 幅:2466px
- 高さ:200px
- ファイルサイズ:17.4KB
- 幅:2460px
- 高さ:160px
- ファイルサイズ:14.5KB
- 幅:2486px
- 高さ:173px
- ファイルサイズ:12.3KB
- 幅:2486px
- 高さ:167px
- ファイルサイズ:15KB
- 幅:1053px
- 高さ:233px
- ファイルサイズ:10KB
- 幅:2533px
- 高さ:273px
- ファイルサイズ:19.2KB
- 幅:2293px
- 高さ:907px
- ファイルサイズ:53.2KB
- 幅:2486px
- 高さ:193px
- ファイルサイズ:17.3KB
- 幅:2473px
- 高さ:207px
- ファイルサイズ:17.1KB
- 幅:2479px
- 高さ:207px
- ファイルサイズ:13.5KB
- 幅:2493px
- 高さ:167px
- ファイルサイズ:14.5KB
- 幅:2470px
- 高さ:202px
- ファイルサイズ:13.5KB
- 幅:2480px
- 高さ:207px
- ファイルサイズ:15.5KB
- 幅:938px
- 高さ:190px
- ファイルサイズ:9.6KB
- 幅:234px
- 高さ:100px
- ファイルサイズ:2.8KB
- 幅:2466px
- 高さ:260px
- ファイルサイズ:22.5KB
- 幅:260px
- 高さ:86px
- ファイルサイズ:3.5KB
- 幅:2481px
- 高さ:192px
- ファイルサイズ:17.7KB
- 幅:2453px
- 高さ:206px
- ファイルサイズ:18.6KB
- 幅:2492px
- 高さ:222px
- ファイルサイズ:15.9KB
- 幅:2493px
- 高さ:220px
- ファイルサイズ:15.7KB
- 幅:2480px
- 高さ:220px
- ファイルサイズ:15.8KB
- 幅:2486px
- 高さ:206px
- ファイルサイズ:15.3KB
- 幅:2480px
- 高さ:140px
- ファイルサイズ:9.2KB
- 幅:2466px
- 高さ:140px
- ファイルサイズ:9KB
- 幅:2486px
- 高さ:146px
- ファイルサイズ:13KB
- 幅:2453px
- 高さ:140px
- ファイルサイズ:12.1KB
- 幅:893px
- 高さ:153px
- ファイルサイズ:8.2KB
- 幅:2479px
- 高さ:220px
- ファイルサイズ:19.2KB
- 幅:2486px
- 高さ:200px
- ファイルサイズ:18.8KB
- 幅:607px
- 高さ:167px
- ファイルサイズ:5.7KB
- 幅:1446px
- 高さ:230px
- ファイルサイズ:13KB
- 幅:2406px
- 高さ:193px
- ファイルサイズ:18.4KB
- 幅:2373px
- 高さ:180px
- ファイルサイズ:16KB
- 幅:2346px
- 高さ:166px
- ファイルサイズ:10.6KB
- 幅:2399px
- 高さ:153px
- ファイルサイズ:11.8KB
- 幅:2380px
- 高さ:160px
- ファイルサイズ:11.4KB
- 幅:993px
- 高さ:100px
- ファイルサイズ:9KB
- 幅:907px
- 高さ:98px
- ファイルサイズ:8KB
- 幅:2493px
- 高さ:227px
- ファイルサイズ:17.7KB
- 幅:2500px
- 高さ:233px
- ファイルサイズ:17KB
- 幅:2487px
- 高さ:393px
- ファイルサイズ:30.2KB
- 幅:2520px
- 高さ:146px
- ファイルサイズ:12KB
- 幅:2487px
- 高さ:167px
- ファイルサイズ:12.3KB
- 幅:2480px
- 高さ:153px
- ファイルサイズ:11.2KB
- 幅:720px
- 高さ:186px
- ファイルサイズ:10.7KB
A particular example of long wave propagation
As a simple example to illustrate the foregoing discussion, we assume a canal of constant width, the bottom profile of which is shown in Figure 2. The regions are
式(56) [Fig.107_118_10]
where
[Fig.107_118_11]
This model is well known for its simple properties of the wave propagation (Rayleigh, Theory of Sound §1486,1926).
Fig.2
[Fig.107_119_01]
-A transition of the finite width.
Take the equation for the volume flux (16)and assume b=unity for simplicity.
The wave equation is now written,
式(57) [Fig.107_119_02]
where
式(58) [Fig.107_119_03]
and
[Fig.107_119_04]
The solutions of (57) are
I) 式(59) [Fig.107_119_05]
II) 式(60) [Fig.107_119_06]
III) 式(61) [Fig.107_119_07]
where m^{2} = n^{2}-1/4. We are concerned with the renection of negatively moving waves in region III and consequently in region I we take only one term which represents the waves moving to the negative x direction.
Defining the reflection coefficient as
式(62) [Fig.107_119_08]
we can derive the formula
式(63) [Fig.107_119_09]
where \mu\=x1/x2.
For large values of \eta\, we may approximate
式(64) [Fig.107_119_10]
If m be imaginary, we may write im=m’, and
式(65) [Fig.107_119_11]
From (63), we see that with increasing n corresponding to the gradual transition from (I) to (III),the reflection diminishes until it vanishes when m log\mu\ = \pi\,i. e.,when
式(66) [Fig.107_120_01]
With a Still more gradual transition, the reflection revives, reaches a maximum, vanishes again when m log\mu\ = 2\pi\, and so on.
To see the situation more clearly, we may set
式(67) [Fig.107_120_02]
where \lambda\2 is the wavelength in region lII and l(=x2-x1) is the width of the transition (II). The critical value of 1/ \lambda\2 corresponding to n defined by (66) is then given by
[Fig.107_120_03]
(64) shows that IRI is inversely proportional to l/\lambda\2 apart from the oscillatory factor.
In Figure 3, the reflection coefficient decreases very rapidly as the ratio of the width of the transition to the wave length in deeper water increases from zero to about 0.5. Another fact is that the amount of reflection does not decrease monotonically, but shows small periodic fluctuations with respect to l/\lambda\2, which suggests interference between the two discontinuities in slope of the bottom, somewhat analogous to the more familiar interactions between discrete interface. A similar phenomenon is found in many cases of the transition of the finite width as in the case of the linear variation of a medium connecting the uniform medium on both sides (Albini and Jahn,1961).
Returning to the theory of wave propagation discussed in the previous section, we may write the solution of (49) corresponding to (59) to (61) as follows:
式(68) [Fig.107_121_01]
式(69) [Fig.107_121_02]
式(70) [Fig.107_121_03]
where
[Fig.107_121_04]
and
[Fig.107_121_05]
Setting [Fig.107_121_06] ,the equation for R is determined from (68) to (70) with the ai of the boundary condirions at \xi\=0 and \xi\ = \xi\1 ,such that
式(71) [Fig.107_121_07]
and
式(72) [Fig.107_121_08]
This is of course the same expression as (63) since、\epsilon\^{2} = 1/4n^{2}. Now, to see the advantage of the expressions (68) to (70) over (59) to (61),(71) is expanded in powers of \epsilon\ assuming a small value for \epsilon\. The first few terms in the expansion are
式(73) [Fig.107_121_09]
where the first term represents the effect of the one-time reflection, and the second term which is of order \epsilon\^{3} gives the effect of the wave reflected three times, and so on.
The reflection coefficient due to one-time reflection only is reduced to
式(74) [Fig.107_121_10]
(74) agrees with (64) and clearly shows that, for weak reflection, the one-time reflection makes the most important contribution to the reflection as a whole. (64) can also be obtained from (32) and indirectly proves that the formula (32) for weak reflection can be derived from the consideration of the one-time reflection only.
If we remember that \xi\1=2\pi\l/\lambda\ (\lambda\ is ghe representative wave length in the transition zone), (74) shows that lRl is maximum at l=(n-1/2)\lambda\/2 and minimum at l=n\lambda\/2 (n=1,2,.…) as a result of the interference of the reflected wave.
- 幅:2660px
- 高さ:366px
- ファイルサイズ:25.3KB
- 幅:940px
- 高さ:140px
- ファイルサイズ:7.9KB
- 幅:2499px
- 高さ:986px
- ファイルサイズ:69.9KB
- 幅:2486px
- 高さ:174px
- ファイルサイズ:12.1KB
- 幅:2633px
- 高さ:560px
- ファイルサイズ:44.8KB
- 幅:413px
- 高さ:207px
- ファイルサイズ:4.8KB
- 幅:2613px
- 高さ:159px
- ファイルサイズ:11.2KB
- 幅:2633px
- 高さ:160px
- ファイルサイズ:13.8KB
- 幅:2619px
- 高さ:160px
- ファイルサイズ:13.2KB
- 幅:2493px
- 高さ:166px
- ファイルサイズ:12.4KB
- 幅:2493px
- 高さ:233px
- ファイルサイズ:20.4KB
- 幅:2500px
- 高さ:193px
- ファイルサイズ:15.3KB
- 幅:2499px
- 高さ:246px
- ファイルサイズ:22.4KB
- 幅:2500px
- 高さ:194px
- ファイルサイズ:14.9KB
- 幅:2462px
- 高さ:168px
- ファイルサイズ:12.8KB
- 幅:1040px
- 高さ:206px
- ファイルサイズ:11.8KB
- 幅:3233px
- 高さ:2579px
- ファイルサイズ:218.9KB
- 幅:2587px
- 高さ:146px
- ファイルサイズ:12KB
- 幅:2613px
- 高さ:167px
- ファイルサイズ:15.3KB
- 幅:2639px
- 高さ:160px
- ファイルサイズ:12.1KB
- 幅:1459px
- 高さ:260px
- ファイルサイズ:12.9KB
- 幅:1440px
- 高さ:173px
- ファイルサイズ:10.2KB
- 幅:450px
- 高さ:142px
- ファイルサイズ:5.4KB
- 幅:2347px
- 高さ:306px
- ファイルサイズ:29.9KB
- 幅:2406px
- 高さ:260px
- ファイルサイズ:23.4KB
- 幅:2373px
- 高さ:260px
- ファイルサイズ:22.3KB
- 幅:2459px
- 高さ:160px
- ファイルサイズ:10.3KB
An example of continuously varying depth
Now, we treat the case where the wave equation can be solved exactly for continuously varying depth. Epstein (1930) and Yoshida (1949) discussed the partial reflection of waves where the wave equation can be transformed into a hypergeometric differential equation, and gave simple expressions for the reflection and transmission coefficients.
The bottom profile is assumed to be
式(75) [Fig.107_121_11]
where y=e
If x varies from -\infty\ to \infty\, h(x) varies from h1 to h2. If 1/lh3I < I1/h1 -1/h2I ,the variation of h(x) is monotonic; otherwise h(x) first rises or falls to a maximum or a minimum and goes back to the final value (see Fig.8a, b) according to h3=0 or h3 < 0 reopectively. For h3 < 0, we restrict the range of h3 to [Fig.107_121_12] lh3I < h1h2/lh1-h2I,because h3 outside of the above condition is of no practical interest in the wave problem. If h3→\infty\, we have
式(76) [Fig.107_121_13]
and the transition is approximately antisymmetric about x=0 as shown in Figure 3c. If h1 = h2,we have
式(77) [Fig.107_122_01]
and we have roughly a symmetric ridge or a valley in the distribution of h according to h3 > 0, or h3 < -h2/4
The reflection coefficient of the wave travelling in the negative x direction from x = \infty\ is given by
式(78) [Fig.107_123_01]
where
[Fig.107_123_02]
and
[Fig.107_123_03]
\lambda\1, and \lambda\2 are the wave lengths at the depth h1 and h2 respectiveoy, and l (=2\pi\/n) is the representative width of the transition zone, although actually h1 and h2 are reached at x= -\infty\ and x=\infty\, respectively.
For h3 > 0 or n^{2}g Ih3I/\omega^{2}\ > 4, \delta\ is always imaginary and for n^{2}g Ih3I/\omega^{2}\ < 4, \delta\ is real. h3 > 0 corresponds to a bank and n^{2}g Ih3I/\omega^{2}\ > 4 corresponds to a trench which is either shallow or narrow, or to the waves of relatively long periods. On the other hand, n^{2}g Ih3I/\omega^{2}\ < 4 corresponds to the case of a deep or broad trench or to the waves of relatively short periods.
It is easily found that if n^{2}g Ih3I/\omega^{2}\ → \infty\, \delta\→i/2 and the reflection coefficient is reduced to
式(79) [Fig.107_123_04]
and further,if p and q are small,we have
式(80) [Fig.107_123_05]
which is in agreement with the well known formula (25) for the strong reflection.
On the other hand, if the transition is very gradual, so that p and q are large, we may approximate (79) by
式(81) [Fig.107_123_06]
The variation of the reflection coefficient with respect to l/ \lambda\2 (=q) is shown in Figure 5.
As is alreday shown for the case of a transition with the finite width (Fig.3),the decrease of IRI is very steep for small values of l/ \lambda\2 and IRI in Figures 3 and 5 are substantially the same. In practice,wemay take l/ \lambda\2 < 1/20 for the applicable range of the formula (25) for strong reflection. For large values of l/ \lambda\2,however,it should be noticed that the reflection coefficient for the case of an infinite transition (Fig.5) caries as exp (2\pi\l/ \lambda\2) and shows no periodic fluctuaion with respect to l/ \lambda\2,in contrast to the case of a transition with finite width (Fig.3) where IRI varies only as (2 \pi\l/ \lambda\2)^{-1} and fluctuates with the period △l = \lambda\/2 (\lambda\2 is the representative wave length of the long waves in the transition zone).The large difference of the behavior of R for large values of l/ \lambda\2 has already been noticed by Schelkunoff (1951) who explaine it as follows:In a continuous model, the infinite regions (-\infty\ , -1/2) and (1/2, \infty\) have an increasingly pronounced effect on the reflection coefficient at x=\infty\ as l/\lambda\2 becomes large. In other words, the various sections of the medium become increasingly better 「matched」to each other and reflections are greatly reduced. It may be said that the fluctuation of R with respect to l/\lambda\2 for the caseof a finite transition is a result of interference of reflected waves, and for the case of a continuous, infinite transition, this interference of reflected waves extends over the whole region and, as a consequence, the resultant reflection coefficient R at x=\infty\ shows no periodic fluctuation but the value itself is greatly reduced.
- 幅:2487px
- 高さ:206px
- ファイルサイズ:18.8KB
- 幅:627px
- 高さ:87px
- ファイルサイズ:6KB
- 幅:2479px
- 高さ:207px
- ファイルサイズ:18.2KB
- 幅:2506px
- 高さ:233px
- ファイルサイズ:17.5KB
- 幅:2493px
- 高さ:4044px
- ファイルサイズ:242.8KB
- 幅:2513px
- 高さ:179px
- ファイルサイズ:21.5KB
- 幅:800px
- 高さ:387px
- ファイルサイズ:15.6KB
- 幅:820px
- 高さ:293px
- ファイルサイズ:11KB
- 幅:2486px
- 高さ:206px
- ファイルサイズ:17.1KB
- 幅:2478px
- 高さ:236px
- ファイルサイズ:20.2KB
- 幅:2460px
- 高さ:187px
- ファイルサイズ:16.7KB
Reflection of waves in the range of intermediate wave lengths
The equation of water waves for the intermediate and short wave lengths does not conform to the wave equation discussed in the preceding sections. The problem of partial reflection and transmission in this range of wave lengths should be treated either by means of a solution of the boundary problem of a potential theory (1) to (3) or a solution derived from the original equation of motion retaining the vertical acceleration of the water particles. The process involved to solve the problem is complicated but, in recent years, the wave on a uniformly sloping beach has been solved completely as far as the theory of waves of small amplitude is concerned and it is shown that for periodic waves, the solution is uniquely determined by assuming a progressive wave moving towards shore with a certain amplitude at infinity (Stoker, Water Waves, Chap.5,1957).
Although the actual breaking of waves near shore is a somewhat different matter, the solution given by the above formulation of the problem shows that the wave energy transported toward shore is exclusively absorbed at the edge of the beach and allows no partial reflection of waves back to infinity. If we allow the partial reflection of wave energy, the problem cannot be determined uniquely because we can choose reflected waves arbitrarily and the singularity of the solution at the edge of the beach may take care of the rest. Thus the question of the partial reflection is not settled by this sort of a solution. As for the partial reflection and transmission of waves, complete solutions are given for surface waves in deep water passing over an infinitely thin barrier (Dean, 1945;Ursell,1947;John,1948) and for waves through a submerged barrier in water of constant depth (Heins,1950), The finite width of the barrier is treated by Takano (1960), but the results are not satisfactory in the sense that the series representing the solution converge very slowly for waves of interest. The problem of wave propagation for a model of continuously varying depth which is similar to (76) has been solved for an arbitrary wave length by Roseau (1952). The outline of his theory is given below.
By means of the transformation*
式(82) [Fig.107_124_02]
*In this section \xi\,\eta\,and \zeta\ are defined differently from the previous sections.
with
式(83) [Fig.107_125_01]
The bottom profile is defined by
式(84) [Fig.107_125_02]
or
式(85) [Fig.107_125_03]
式(86) [Fig.107_125_04]
where \alpha\ is a constant. The geometry of the profile is shown in Figure 6,and the relation between the angle H at the maximum slope and \alpha\ is
式(87) [Fig.107_125_05]
where h1 and h2 are the asymptotic depth of water for negative and positive infinities respectively.
We can also define the representative width of the transition by putting \xi\ =+-\pi\. For this definition the length of the transition is given by
式(88) [Fig.107_125_07]
and
式(89) [Fig.107_125_08]
and the totallength of the transition zone is
式(90) [Fig.107_125_09]
If we define the representative slope by
式(91) [Fig.107_125_10]
the representative slope is found to be proportional to \alpha\
For this model of the bottom profile, the set of equations (1) to (3) in terms of \phi\(\omega\) can be transformed into the new set of equations. These equations are, in terms of new velocity potential \varphi\(\zeta\), given by
式(92) [Fig.107_125_11]
式(93) [Fig.107_125_12]
and
式(94) [Fig.107_126_01]
where
[Fig.107_126_02]
The solution of the above set of equations is sought in the form:
式(95) [Fig.107_126_03]
where H(u) and K(u) are analytic function of u, C and P the contours in the u-space, and l,m, p, and q are constants, the relations of which are determined later. If the integral is convergent, it is seen that \varphi\(\xi\,\eta\) satisfies the set of equations. The functional characteristics of H(u) and K(u) are investigated in detail and it is shown that
式(96) [Fig.107_126_04]
where +- \mu\ and +- \lambda\ are simple poles of H(u) and K(u) respectively and are given as the real roots of the equation
式(97) [Fig.107_126_05]
when h is set equal to h1 and h2 respectively.
Finally,it is shown that the asymptotic form of the solution (95) is reduced to
式(98) [Fig.107_126_06]
for \xi\ →+\infty\ ,and
式(99) [Fig.107_126_07]
for \xi\ →-\infty\
Now assume a progressive wave for \xi\ →-\infty\ ,so that in (99) we set
式(100) [Fig.107_126_08]
Then from (98) the reflection coefficient is expressed by
式(101) [Fig.107_126_09]
where the relations between l,m, p, q, H, and K are used.
It is noticed that this formula is quite similar in form to the reflection formula (79) given for long waves. In fact, for long waves,(97) may be approximated by
式(102) [Fig.107_126_10]
so that
式(103) [Fig.107_126_11]
and
式(104) [Fig.107_126_12]
where [Fig.107_127_01] and [Fig.107_127_02] are the wave numbers for\xi\→ -\infty\ and \xi\→\infty\ respectively. Making use of (88) and (89), the equations may be rewritten
式(105) [Fig.107_127_03]
Thus, we have
式(106) [Fig.107_127_04]
(106) is analogous to (79).The only difference is that in the present model l1 \neq\ l2.
The bahevior of the reflection coefficient with respect to 1/ \lambda\2, where l=l1+l2,is almost the same as the case of the former example, namely, for long waves travelling over a bottom of relatively steep slope, the reflection coefficient is reduced to the familiar formula
式(107) [Fig.107_127_05]
and for a bottom of very gradual slope, the reflection ill be
式(108) [Fig.107_127_06]
The general behavior can, therefore, be estimated from Figure 5. It is noticed that in this case, too, no particular intensification of the reflection due to the interference is found.
The reflection coefficient is shown in Figure 7 as a function of x,for various values of \omega^{2}\h/g and h1/h2. Fr.m Figure 7 it is seen that the variation.f the reflection coefficient with respect to x is larger for an intermediate wave than for long waves, at least in the range of large x, as is expected.
As a particular example, assume a model of a continental slope as follows:
l=100km and 50 km
h1=200 m, h2=4000m ; h1/h2=0.05
From(91),x is given by
[Fig.107_127_07]
For this model, the reflection coefficients R corresponding to particular values of \omega^{2}\h/g or period are shown in Table 1.
if we consider that, for the abrupt change of the depth from h2 to h1,the reflection coefflcient is 0.635, the reduction of the coefficient is considerable,in particular for the waye of 10 minutes in period or smaller.
The Roseau’s theory is valid for \alpha\ < \pi\ only. Therefore, the limiting procedure does not yield a reflection coefficient for the abrupt change of depth, although practically, the value for the long waves at \alpha\ \pi\ is almost identical to the value expected from the ordinary theory. As is seen in Figure 7, the reflection coefficient R for intermediate waves is increasing with \alpha\ even at \alpha\ \pi\ , and, therefore, the limiting value for the abrupt change of the medium would be a little higher than the values given for \alpha\ = \pi\.
The situation may be examined by a method which makes use of the series form of the solution (see Takano,1960). The solution of the potential theory can be expressed by a series, each term of which satisfies the surface and bottom boundary conditions. Then, the boundary condition at the junction of two different depths can be satisfied by a suitable choice of the constant for each terms. The inspection of these series reveals that the particular effect of the boundary extends only to a distance comparable to the depth, because, except for the first term, the remaining terms all die out exponentially as e^{-knx} where kn is of order n\pi\/h. The detailed discussion is not attempted here, but it can be shown that the ordinary reflection formula for long waves (25) is the first approximation to the complete formula for reflection neglecting all terms involving exponential decay term in the horizontal direction.
- 幅:3179px
- 高さ:2599px
- ファイルサイズ:217KB
- 幅:2486px
- 高さ:220px
- ファイルサイズ:16.4KB
- 幅:2513px
- 高さ:147px
- ファイルサイズ:11.7KB
- 幅:2466px
- 高さ:220px
- ファイルサイズ:19.4KB
- 幅:2513px
- 高さ:233px
- ファイルサイズ:20.6KB
- 幅:2507px
- 高さ:233px
- ファイルサイズ:20.8KB
- 幅:2472px
- 高さ:265px
- ファイルサイズ:23.4KB
- 幅:2393px
- 高さ:980px
- ファイルサイズ:65KB
- 幅:2493px
- 高さ:200px
- ファイルサイズ:14.2KB
- 幅:2499px
- 高さ:200px
- ファイルサイズ:14.3KB
- 幅:2486px
- 高さ:180px
- ファイルサイズ:13.5KB
- 幅:2509px
- 高さ:213px
- ファイルサイズ:18.1KB
- 幅:2366px
- 高さ:160px
- ファイルサイズ:11KB
- 幅:2367px
- 高さ:220px
- ファイルサイズ:15KB
- 幅:2326px
- 高さ:107px
- ファイルサイズ:8.1KB
- 幅:1853px
- 高さ:233px
- ファイルサイズ:15.8KB
- 幅:2689px
- 高さ:610px
- ファイルサイズ:64.9KB
- 幅:2459px
- 高さ:227px
- ファイルサイズ:29.2KB
- 幅:2513px
- 高さ:206px
- ファイルサイズ:14.8KB
- 幅:2666px
- 高さ:433px
- ファイルサイズ:56.9KB
- 幅:2666px
- 高さ:427px
- ファイルサイズ:54KB
- 幅:2532px
- 高さ:200px
- ファイルサイズ:16.2KB
- 幅:2506px
- 高さ:240px
- ファイルサイズ:18.6KB
- 幅:2506px
- 高さ:193px
- ファイルサイズ:13.8KB
- 幅:2520px
- 高さ:200px
- ファイルサイズ:13.2KB
- 幅:2499px
- 高さ:173px
- ファイルサイズ:11.3KB
- 幅:477px
- 高さ:227px
- ファイルサイズ:5.9KB
- 幅:493px
- 高さ:220px
- ファイルサイズ:6KB
- 幅:2432px
- 高さ:173px
- ファイルサイズ:12.2KB
- 幅:2492px
- 高さ:393px
- ファイルサイズ:32KB
- 幅:2506px
- 高さ:279px
- ファイルサイズ:20.1KB
- 幅:2503px
- 高さ:222px
- ファイルサイズ:22.2KB
- 幅:1086px
- 高さ:187px
- ファイルサイズ:10.4KB
- 幅:1767px
- 高さ:713px
- ファイルサイズ:58.2KB
- 幅:px
- 高さ:px
- ファイルサイズ: bytes
The two dimensional propagation of long waves
From the equations for long waves (8) , (9), and (10), we may derive the twodimensional wave equation in the following form:
式(109) [Fig.107_129_01]
If we consider the weak reflection only, and assume that the horizontal scale to of all the inhomogeneities in the spatial distribution of the depth are greater than the wave length \lambda\,or more precisely, \lambda\/l0 << l then the propagation of waves in an inhomogeneous medium is described by the equation
式(110) [Fig.107_129_02]
where \eta\ is taken to be periodic in time (~e^{i\omega\t}), and [Fig.107_129_03]
Setting [Fig.107_129_04]
we can derive the equations
式(111) [Fig.107_129_05]
and
式(112) [Fig.107_129_06]
We note that ▽S is of order k and the wave amplitude A can change appreciably only in distances of the order of the dimensions of the inhomoneneities in the medium l0.
Therefore,▽^{2}A/A is of order no greater than l/l0^{2}. lf we assume \lambda\ << l0, the first two terms in (111) are dominant, and the amplitude term can be neglected, so that we have
式(113) [Fig.107_129_07]
The equations (112) and (113) form the equations of geometric optics, and if the variation of depth within one wave length is small compared with the total depth and if the change of wave amplitude is negligible between neighboring rays, the simple laws of refraction and shoaling apply and the wave height follows Green’s law just as if the neighboring two rays constituted solid lateral boundaries.
However, even if the condition \lambda\ << l0 is met, the solution obtained from the equations of geometric optics becomes unsuitable for distances l that exceed the critical distance Lcr=l0^{2}/\lambda\. At such large distances, we can no longer neglect the diffraction of the wave by depth inhomogeneities, regardless of the smallness of the diffraction angle \theta\〜 \lambda\/1. Because, at a distance L from the obstacle, the size of the diffracted bundle will be of order \theta\L〜 \lambda\L/l,and for the geometric shadow of the obstacle not to be appreciably changed, it is necessary for the relation \lambda\l/l < 1 to hold. In order to take into account diffraction effects, it is necessary to start from the wave equation.
The following treatment is essentially due to Tatarski (1961). Write (110) in the form
式(114) [Fig.107_129_08]
where [Fig.107_130_01]
and we call n the refractive index.
Set
式(115) [Fig.107_130_02]
Then we have
式(116) [Fig.107_130_03]
Set \phi\ = \phi\0+\phi\1 ; \phi\0 satisfies the equation
式(117) [Fig.107_130_04]
Substituting \phi\ = \phi\0+\phi\1 in (116) ,we have
式(118) [Fig.107_130_05]
In the case where l▽\phi\1I <= n1I▽\phi\0I, we can neglect the term.(▽\phi\^{2})
Finally we obtain the equation
式(119) [Fig.107_130_06]
which is valid when the condition
[Fig.107_130_07]
are met. Since I▽\phi0\l k=2\pi\/\lambda\ ,the second condition can be written in the form
[Fig.107_130_08]
and expresses the smallness of the change of \phi\1 over distances of the order of a wavelength. By using the substitution [Fig.107_130_09] (119) can be reduced tothe form
式(120) [Fig.107_130_10]
Since [Fig.107_130_11] ,the above equation is equivalent to the perturbation equation derived from (110) by assuming \xi\ = \xi\0+\xi\1 so that w=\xi1\ and [Fig.107_130_12].
We now find expressions for the amplitude and phase fluctuations of the wave.Since
[Fig.107_130_13]
we have
[Fig.107_130_14]
Therefore, we may write
式(121) [Fig.107_130_15]
and
式(122) [Fig.107_130_16]
The validity of these expression is given by the conditions \lambda\I▽S1I << 2\pi\ and \lambda\I▽xI << 2\pi\ rather than requiring the smallness of the perturbations x and S1 themselves.Since the inequality Ixl < IS1I is usually satisfied, the method is valid for the conditions
式(123) [Fig.107_130_17]
As is well known, the solution of (120) has the form
式(124) [Fig.107_130_18]
and for the quantity [Fig.107_130_19] ,we obtain
式(125) [Fig.107_130_20]
where G is the free-space Green’s function,\Sigma\ is the area, and dS’ is the two-dimensional element. ,
Since [Fig.107_131_01] satisfies the equation
式(126) [Fig.107_131_02]
which is equivalent to (117)
Now G is given by
式(127) [Fig.107_131_03]
where H is a Hankel function of the first kind.
Sett’ng
式(128) [Fig.107_131_04]
(125) may be written formany;
式(129) [Fig.107_131_05]
In general, a stationary random function can be expressed in the form of a stochastic Fouries-Stieltjes integral with random complex amplitudes. Thus, we may write
式(130) [Fig.107_131_06]
and
式(131) [Fig.107_131_07]
and (129) becomes
式(132) [Fig.107_131_08]
Multiply by it [Fig.107_131_09] and integrate with respect to y. Taking the relation of the Dirac delta function
式(133) [Fig.107_131_10]
into consideration, the left hand side of (132) is t ransformed into
[Fig.107_131_11]
Thus, we have
式(134) [Fig.107_131_12]
Changing the order of integration and considering that the main effect observed at the point of observation comes from the region between the original source of waves and the point of interest, we have
式(135) [Fig.107_131_13]
For Plane monochromatic waves, we have [Fig.107_131_14], so that (128) gives
式(136) [Fig.107_131_15]
Making use of the formula
式(137) [Fig.107_131_16]
we can deduce the relation
式(138) [Fig.107_132_01]
This formula shows that the field inhomogeneities characterized by the wave number k2 (i,e., by the dimension 1=2 \pi\/k2) owe their origin to inhomogeneities of the medium with the same characteristic wave number k2.
Since x=Re \phi\1 and S1=Im\phi\1, we have
式(139) [Fig.107_132_02]
and
式(140) [Fig.107_132_03]
where an asterisk indicates the complex conjugate quantity. Denoting the spectral amplitudes of the random fields x(r1) and S1(r1) by da(k2,L),and \sigma\(k2,L), we may write from (138)
式(141) [Fig.107_132_04]
and
式(142) [Fig.107_132_05]
Since we have assumed k2 << k or \lambda\ << l,we may approximate [Fig.107_132_06], and (141) and (142) will reduce to (7.37) and (7.38) of Tatarski’s book (1961),exactly as is expected.The refractive index inhomogeneities with dimensions l which are located at a distance (L-x’) from the observation point appear with weight sin (\pi\\Lambda\^{2}/l^{2}), where \Lambda^{2}\= \lambda\(L-x’) is the square of the radius of the first Fresnel zone.
To obtain the relations between the spectral densities of the correlation functions, we take the following averages:
式(143) [Fig.107_132_07]
But, according to the general foxmula, we have
式(144) [Fig.107_132_08]
and
式(145) [Fig.107_132_09]
where Fn(k2,x’-x’’) is the two-dimensional spectral function of the refractive index and FA(k2,0) is the one-dimensional spectral denssity of the correelation function of the fluctuation of x in the plane x=constant.
Substituting (144) and (145) into (143),we obtain
式(146) [Fig.107_132_10]
We introduce new variables \xi\=x’-x" and 2\eta\ =x’+x". Since Fn does not depend on \eta\, the integration of (146) with respect to \eta\ can be carried out explictly,and yields
式(147) [Fig.107_132_11]
Now since Fn(k2,\xi\) falls off very rapidly to zero for k2\xi\>~ 1,we can assume k2^{2}\xi\/(2k) << 1 in the important region of integration of (147). Moreover,we are interested in values of k2 which satisfy the condition 2\pi\/k2 << L, we may consider the region \xi\ << L only. Taking these simplifications into account,we have
式(148) [Fig.107_133_01]
Here, we may set
式(149) [Fig.107_133_02]
and we may replace L by \infty\ in (148) because of the special nature of Fn(k2,\xi\). Therefore, we have
式(150) [Fig.107_133_03]
In the same way, we can derive for the phase fluctuation,
式(151) [Fig.107_133_04]
Equations (150) and (151) relate the one-dimensional special densities of the correlation functions of the amplitude and phase fluctuations of the wave on the line x=L to the twodimensional spectral density of the correlation function of the refractive index.
It follows from (150) and (151) that the phase fluctuations are always larger than the fluctuations of logarithmic amplitude. Taking the relation k=2\pi\/\lambda\ into consideration、 we can discuss the following three cases separately, where l0 and L0 are the inner and outer scales of the inhomgeneities of the medium.
In the case where, [Fig.107_133_05], the spectrum of the correlation function of the amplitude fluctuations is concentrated near the point 2\pi\/l0. It follows from general properties of the Fourier transform that the correlation function has a characteristic scale of order lo. The amplitude fluctuations of the wave do not depend on the frequency of the wave but grow with the distance as L^{3}, while the phase fluctuations are proportional to the square of the frequency and to the distance L.
In the case where [Fig.107_133_06], the spectrum of the correlation function of the amplitude fluctuations is concentrated near the point [Fig.107_133_07]. It follows from this that the correlation function of the amplitude fluctuations on the Line x=L has a characteristic scale of order [Fig.107_133_08]. The form of the correlation function of the amplitude fluctuations depends on the concrete form of the spectral density \Phi\n(k2) of the refractive index fuuctuations.
In the case where、[Fig.107_133_09], the field of spectral density\Phi\n(k2) is not homogeneous nor isotropic for k2 < 2 \pi\/L0. Therefore, FA and FS have me aning only in the region k2 >> 2 \pi\/L0,where
[Fig.107_133_10]
Thus, FA and FS are equal to one another and proportional to the square of the frequency and to the distance L traversed by the wave. However, the chief contribution to the spectrum of the correlation function of the amplitude fluctuations is made by large scale inhomogeneities in the interval ([Fig.107_133_11]) and it is not possible to discuss the behavior of the spectrum within the present formulation of the problem.
In all the cases considered, the largest wave numbers which participate in the spectral expansions of the amplitude and phase fluctuations of the wave are of order 2 \pi\/lo. It follows from this that the correlation functions of the amplitude and phase fluctuations of the wave change quite slowly over a distance of order lo.
For cylindrical waves diverging from a point source.,the similar procedure can
be applicable. Now, we have in (128)
式(152) [Fig.107_133_12]
Making use of the asymptotic form for the Hankel functions, we may transform (152) into
式(153) [Fig.107_134_01]
where it is assumed that x,x’and x-x’ are generally large compared with y, y’, and y-y’ because the main contribution of the waves comes from the area with the angle \lambda\/lo at the point of interest and its axis on x.
The manipulations similar to the case of plane waves yields
式(154) [Fig.107_134_02]
Field inhomogeneities characterized by the wave number k2 owe their origin to inhomogeneities of the medium with characteristic wave number k2L/x, or with dimension l’=lx’/L. These inhomogeneities are at the distance x’ from the wave source. The factor x’/L takes into account the magnification of the dimensions of the image due to illumination by a divergent ray bundle.
It is possible to discuss the spectral densities of the correlation functions of the amplitude and phase fluctuations in relation to the two-dimensional spectral density of the correlation function of the refract ive index (inhomogeneities of the medium), by means of a method similar to that for the case of plane waves. However, because of the limited application of the present model to the actual problem of tsunamis in an enclosed ocean, no further discussion is attempted here (see Tatarski, Chap.9,1961).
- 幅:2559px
- 高さ:186px
- ファイルサイズ:17.1KB
- 幅:2533px
- 高さ:199px
- ファイルサイズ:14.8KB
- 幅:1407px
- 高さ:233px
- ファイルサイズ:16.5KB
- 幅:800px
- 高さ:147px
- ファイルサイズ:8.6KB
- 幅:2506px
- 高さ:180px
- ファイルサイズ:13.6KB
- 幅:2546px
- 高さ:133px
- ファイルサイズ:10.4KB
- 幅:2520px
- 高さ:146px
- ファイルサイズ:10.9KB
- 幅:2512px
- 高さ:160px
- ファイルサイズ:12.9KB
- 幅:2373px
- 高さ:220px
- ファイルサイズ:19.3KB
- 幅:2526px
- 高さ:146px
- ファイルサイズ:12.3KB
- 幅:2526px
- 高さ:140px
- ファイルサイズ:12.4KB
- 幅:2532px
- 高さ:147px
- ファイルサイズ:12.8KB
- 幅:2540px
- 高さ:194px
- ファイルサイズ:18.8KB
- 幅:2532px
- 高さ:173px
- ファイルサイズ:14.1KB
- 幅:1287px
- 高さ:167px
- ファイルサイズ:10.5KB
- 幅:651px
- 高さ:121px
- ファイルサイズ:4.7KB
- 幅:404px
- 高さ:97px
- ファイルサイズ:3.5KB
- 幅:2507px
- 高さ:150px
- ファイルサイズ:13.2KB
- 幅:313px
- 高さ:133px
- ファイルサイズ:3.5KB
- 幅:705px
- 高さ:105px
- ファイルサイズ:6.9KB
- 幅:1259px
- 高さ:133px
- ファイルサイズ:10.8KB
- 幅:1446px
- 高さ:146px
- ファイルサイズ:11.8KB
- 幅:2373px
- 高さ:200px
- ファイルサイズ:17KB
- 幅:2440px
- 高さ:166px
- ファイルサイズ:15.1KB
- 幅:2520px
- 高さ:147px
- ファイルサイズ:13.3KB
- 幅:2419px
- 高さ:193px
- ファイルサイズ:15.9KB
- 幅:424px
- 高さ:200px
- ファイルサイズ:7.2KB
- 幅:2487px
- 高さ:240px
- ファイルサイズ:21.2KB
- 幅:461px
- 高さ:130px
- ファイルサイズ:4.9KB
- 幅:620px
- 高さ:166px
- ファイルサイズ:6.4KB
- 幅:945px
- 高さ:221px
- ファイルサイズ:9.8KB
- 幅:2399px
- 高さ:240px
- ファイルサイズ:18.9KB
- 幅:2420px
- 高さ:219px
- ファイルサイズ:18.1KB
- 幅:2406px
- 高さ:247px
- ファイルサイズ:18.4KB
- 幅:2460px
- 高さ:219px
- ファイルサイズ:17KB
- 幅:2352px
- 高さ:218px
- ファイルサイズ:22.7KB
- 幅:400px
- 高さ:167px
- ファイルサイズ:4.6KB
- 幅:2440px
- 高さ:213px
- ファイルサイズ:15.1KB
- 幅:1327px
- 高さ:260px
- ファイルサイズ:14.7KB
- 幅:2513px
- 高さ:240px
- ファイルサイズ:24.2KB
- 幅:2540px
- 高さ:247px
- ファイルサイズ:27.6KB
- 幅:329px
- 高さ:139px
- ファイルサイズ:3KB
- 幅:2500px
- 高さ:213px
- ファイルサイズ:17.7KB
- 幅:2547px
- 高さ:340px
- ファイルサイズ:32.2KB
- 幅:2526px
- 高さ:306px
- ファイルサイズ:29.6KB
- 幅:2553px
- 高さ:227px
- ファイルサイズ:20.8KB
- 幅:2519px
- 高さ:233px
- ファイルサイズ:20.9KB
- 幅:2566px
- 高さ:500px
- ファイルサイズ:45KB
- 幅:2533px
- 高さ:320px
- ファイルサイズ:29.6KB
- 幅:873px
- 高さ:213px
- ファイルサイズ:10.3KB
- 幅:2613px
- 高さ:400px
- ファイルサイズ:39.9KB
- 幅:2593px
- 高さ:206px
- ファイルサイズ:21.9KB
- 幅:2613px
- 高さ:173px
- ファイルサイズ:21KB
- 幅:2536px
- 高さ:241px
- ファイルサイズ:27.9KB
- 幅:2553px
- 高さ:413px
- ファイルサイズ:39.1KB
- 幅:2520px
- 高さ:260px
- ファイルサイズ:23.6KB
- 幅:2515px
- 高さ:190px
- ファイルサイズ:15KB
- 幅:2533px
- 高さ:293px
- ファイルサイズ:24.6KB
- 幅:2501px
- 高さ:265px
- ファイルサイズ:22.7KB
- 幅:382px
- 高さ:89px
- ファイルサイズ:3.2KB
- 幅:599px
- 高さ:90px
- ファイルサイズ:4.9KB
- 幅:320px
- 高さ:93px
- ファイルサイズ:3.3KB
- 幅:213px
- 高さ:107px
- ファイルサイズ:2.1KB
- 幅:370px
- 高さ:111px
- ファイルサイズ:3.6KB
- 幅:1307px
- 高さ:206px
- ファイルサイズ:12.5KB
- 幅:302px
- 高さ:84px
- ファイルサイズ:2.7KB
- 幅:2573px
- 高さ:366px
- ファイルサイズ:26.9KB
- 幅:2532px
- 高さ:253px
- ファイルサイズ:27.8KB
- 幅:2500px
- 高さ:273px
- ファイルサイズ:26.3KB
Summary and conclusion
The partial reflection of long waves in an one-dimensional canal of variable cross section is discussed. The basic equations are (11) and (12), Various methods to solve the wave equations (11) and (12) are introduced.(29) is the equation in terms of the reflection coefficient,(39) and (40) are the set of equations showing the mechanism of wave reflection explicitly and defining the local reflection modulus (47). (48) and (49) are the wave equations which include the local reflection modulus explicitly. The method of characteristics applied to a linear wave problem is introduced in (50) and (51),which show the mechanism of reflection clearly just as (39) and (40) do.
In general, the contribution of the one-time reflection of the incident wave is the most important part of the reflected wave as a whole and, for the case of weak reflection, it is enough to consider the one-time reflection (32) only. Moreover, for weak reflection, the transmitted wave is approximated by Green’s formula quite well even though the part of the wave is actually reflected continuously.
Some particular examples are discussed in detail and it is shown that for long waves, the reflection coefficient is approximated by the well known formula (25) for the abrupt change of the depth if l/\lambda\2 is less than, say,one-twentieth. With increasing values of l/ \lambda\2, the reflection coefficient decreases very rapidly and for l/\lambda\2 >> 1, the reflection is very small. The differences of the behavior of the reflection coefficients for large values of l/\lambda\2,in the cases of finite width of a transition and of continuous, infinite width of a transition, are considered to be a consequence of the different interference of the reflected waves in both cases.
The partial reflection of intermediate and short waves is treated, and the reflection coefficients are shown in Figure 7 as a function of l/\lambda\2 for particular values of \omega^{2}\h2g9 and h1/h2. In a model resembling a continental slope and an ocean, the reflection coefficient of waves of 10 minutes in period is considerably smaller than that of waves of 20 minutes in period.
The two-dimensional propagation of long waves over a bottom of variable depth is discussed briefly for the case of weak reflection following the method presented by Tatarski for the electromagnetic wave propagation in a turbulent atmosphere. The relation of the one-dimensional spectral density of the correlation functions of the amplitude and phase fluctuations of the wave on a line x=L to the two-dimensional spectral density of the correlation function of the refractive index defined by n^{2} = 1/(gh) is formulated.
If we consider the problem from the standpoint of the tsunami propagation in an enclosed ocean, the status of research on the two-dimensional propagation of waves is still quite unsatisfactory. The interference of waves due to total reflection at the continental boundary, which seems to be actually taking place in the later stage of the wave train, might be treated, from the statistical point of view, differently from the view introduced in this paper. It may be worthwhile to try numerical method of approach to the problems of tsunami propagation but due care must be taken in the selection of a grid system because we must distinguish between the strung and weak reflection within one grid interval, and. in addition, the grid interval △x must be small compared with the wave length (actually,[Fig.107_135_01] seems to be enough for the case of weak reflection).
- 幅:401px
- 高さ:91px
- ファイルサイズ:2.8KB
References
Albini F.A.,and R. G. Jahn.1961:Reflection and transmission of electromagnetic waves at electron density gradients.
Applied Physics, 32,75-82.
Bremmer, H.,1949:The propagation of electromagnetic waves through a stratified medium and its W-K-B approximation for oblique incidence. Physica,15,593-608.
Bremmer, H.,1951:The W-K-B approximation as the first term of a geometric-optical series. Comm. Pure and Appl. Math.,4,105-115.
Courants R.,and K. O. Friedrichs,1948:Supersonic flow and shock waves. Inter-science Publishers, Inc.,New York.
Dean, W. R.,1945:0n the reflexion of surface waves by a submerged plane barrier.
Proc. Cambridge Phil. Soc.,41,231-238.
Epstein, P. S.,1930:Reflection of waves in an inhomogeneous absorbing medium.
Proc. Nat. Acad. Sci.16,627-637.
Heins. A. E.,1950:Water waves over a channel of finite depth with a submerged plane barrier. Canadian Math.,2, 210-222.
John.F.,1948: Waves in the presence of an inclined barrier. Comm. Pure and Appl.
Math.,1,149-200.
Ogawa, K.,and K. Yoshida.1959: A practical method for a determination of reflection of long gravitational waves. Records Oceanogr. Works in Japan. 5,38-50.
Pierce. J. R.,1943:Note on the transmission line equations in terms of impedance.
Bell Sys. Tech.,22.263-265.
Rayleigh, Lord,1926:Theory of Sound Vo1.1,2nd Ed., MacMillan and Co. Ltd.,
London.
Reid, R.0.,1957:Forced and free surges in a narrow basin of variable depth and width;
A numerica1 approach. Tech. Report Ref.57-25T, Texas A&M Research
Foundation 60 PP.
Roseau, M.,1952:Contribution a la theorie des ondes liquides de gravite en profondeur variable. Publ. Sci. et Tech. du Ministere de 1’Air. N’ 275,90 pp.
Schelkunoff, S. A.,1951:Remarks concerning wave propagation in stratified media,
Comm. Pure and App1. Math.,4.117-128.
Stoker. J. J.,1957:Water waves. Interscience Publishers, Inc., New York.
Takano, K.,1960:Effets d’un obstacle parallelepipedique sur la propagation de la
houle. La Houille Blanche, N’3-1960.247-267.
Tatarski, V.1.,1961:Wave propagation in a turbulent medium. Translated by R. A.
Silverman from the Russian. McGraw-Hill Book Company, Inc., New York.
Ursell, F.,1947:The effect of a fixed vertical barrier on surface waves in deep water. Proc. Cambridge Phil. Soc.,43,374-382.
Ursell,F.,1953:The long-wave paradox in the theory of gravity waves. Proc.
Cambridge Phil. Soc.,49,685-694.
Walker. L. R. and Wax, N.,1946:Non-uniform transmission lines and reflection coefficients. J. Applied Physics, 17,1043-1045.
Yoshida, K.,1948:0n the partial reflection of long waves. Geophys. Notes,Geophys.
Ins.,Tokyo Univ.,N’31,14 pp.
ANote on the Generation of Boundary Waves of Kelvin Type* Kinjiro KAJIURA**
Abstract
The generation of Kelvin waves along a straight boundary due to the reflection of cylindrical waves is studied in a rotating sea of constant depth. The result is applied to the discussion of an internal wave of Kelvin type generated by an atmospheric disturbance. In a two-layer model of the ocean a simple formula is derived which gives the Kelvin wave height in terms of the quantities associated with an atmospheric desturbance. The effect of the wind very close and parallel to the coast is most favorable to excite the internal Kelvin wave travelling along the coast to the right (facing offshore in the northern hemisphere), the amplitude of which at the water surface is of the order of 10 cm for a wind field of reasonable scale.
1. Introduction
It is well known that the Kelvin wave is a phenomenon essentially related to the existence of a lateral boundary on the rotating earth and tidal theorists who studied oscillations of a water body on the earth were well aware that the reflection of long waves at a boundary is seriously complicated by the fact of the earth’s rotation (Lamb, Hydrodynamics pp.307-320,1932). Tidal oscillations in gulfs and rectangular basins on the rotating earth were throughly discussed by G.I. TAYLOR (1922) who could show that a train of Kelvin waves is reflected near the head of a gulf in such a way that wave crests progress counterclockwise (in the northern hemisphere). Recently, J. CREASE (1956) studied the diffraction of simple harmonic plane waves on the rotating earth due to a semi-infinite barrier and fourld that the wave of Kelvin type travels without attenuation to the right (facing offshore) along the shadow side of the barrier.
Later study (CREASE,1958) has shown that the waves propagating into a semi-infinite channel are modified by the rotation of the system to generate the waves of Kelvin and Poincare types.
These studies have clarified the behavior of plane harmonic waves near the boundary and suggest an interesting possibility that the existence of a straight boundary might transform cylindrical harmonic waves diverging from a source region through complicated reflection at the boundary into waves of Kelvin type at a large distance from the source region (to the right facing offshore in the northern hemisphere) where ordinary harmonic plane waves propagating parallel to the straight boundary will be nullified by the reflected waves which are \pi\ out of phase to the incident waves.
Since the effect of the earth’s rotation on simple harmollic waves depends on the magnitude of the parameter f/\omega\ (\omega\ :frequency of the water wave, f:the Coriolis parameter), ordinary free long progressive waves of the barotropic mode such as tsunamis are considered too small in period to be appreciably influenced by the rotation of the earth but, in tidal theories in which f/\omega\ is of order unity, the geostrophic effect is always taken into account because of its importance. The above discussion can equally be applicable to internal long waves (baroclinic waves) the periods of which may be considerably larger than those of barotropic waves of the same wave lengths, so that the effect of the earths rotation on these internal waves of long wave lengths of the order of 100km is expected to be very important.
The detailed discussions of the behavior of barotropic and baroclinic waves generated by an atmospheric disturballce ln a two-layer boundless ocean are already given by VERONIS and STOMMEL (1956), VERONIS (1956), KAJIURA(1958). And ICHIYE(1958) and YOSHIDA (1960) have treated the similar problem in a bounded ocean, in which the boundary waves of the baroclinic mode are already noticed.
In the present paper, the reflection of long water waves, especially of cylindrical waves, at a straight boundary in a semi-infinite sea of constant depth is studied and the result is utilized to investigate the possibility of the Kelvin wave generation by means of an atmospheric disturbance of a finite dimension near the boundary. Unlike the case discussed by YOSHIDA who treated harmonic disturbances only, the present discussion is more general in a sense that the distribution of the atmos- pheric disurbance can be arbitrary in a framework of long wave approximation.
* Rec. Dec.25,1961.
** Earthquake Regearch Institute, University of
Tokyo.
2.Fundamental equations
The equations governing the generation and propagation of long waves of small amplitude in water of uniform density on the rotating earth may be written, subject to certain linearizing approximatlons,
式(1) [Fig.107_137_01]
式(2) [Fig.107_137_02]
and the equation of continuity is
式(3) [Fig.107_137_03]
were t is time, x and y are right-handed horizontal co-ordinates, u and v are volume fluxes per unit sidth in the x and y direction respectively,\zeta\ is the elevation of the water surface from the undisturbed level, d is the constant depth of water, g is the acceleration due to gravity and f is the Coriolis parameter.
Fx and Fy are the components of an external force F of the atmospheric origin in the x and y direction respectively and the frictional force at the bottom is neglected. (see Fig.1).
The initial conditions is assumed to be
式(4) [Fig.107_137_06]
so that no disturbance exists in water before the application of external forces:
式(5) [Fig.107_137_07]
The straight boundary is assumed to lie on the y-axis with x-axis positive in the offshore direction. The boundary condition is given by no net transport of water across the boundary so that
式(6) [Fig.107_137_08]
For a large distance from the source region, waves should be divergent to be consistent with the physical setting of the problem.
From (1) to (3), the equation in terms of \zeta\ is derived as follows
式(7) [Fig.107_137_09]
where
式(8) [Fig.107_137_10]
The corresponding expresslons for the initial and boundary conditions are
式(9) [Fig.107_137_11]
and
[Fig.107_137_12]
where
式(10) [Fig.107_137_13]
*The last condition, [Fig.107_137_04], is derived from the equations of motion with the aid of (4) and (5).
Furthermore, \zeta\ =0 for [Fig.107_138_01] with finite t.
In the derivation of (7), the effect of the variation of the Coriolis parameter f with latitude is neglected which is small if the condition \beta\/2\omega\ << f/c is satisfied where \beta\ is the latitude change of the Coriolis parameter and \omega\ is the representative frequency of the wave considered.
By means of the Green’s function G(x,y,t,x0,y0,t0) defined by
式(11) [Fig.107_138_02]
with the boundary condition,
式(12) [Fig.107_138_03]
the solution of (7) with the conditions (5), (9) and (10) can be constructed formally as follows:
式(13) [Fig.107_138_04]
where \delta\(x) is Dirac’s delta function and dS0, is the surface element and S is taken for the region of the atmospheric disturbance. The term representing the initial conditions vanishes because of the conditions (9) and the last term is related to the boundary condition (10). (MORSE and FESHBACH,1953, p.854).
It may be remarked here that, for a twolayer model of the ocean, waves of the barotropic mode and of the baroclinic mode can be described by the similar equation (7) provided that c^{2} is understood as the minimum phase velocity (or the maximum group velocity) for each mode respectively, and the importance of the Coriolis parameter in wave motion shows up very distinctly for the internal wave because of the large magnitude of f/c baroclinic compared with f/c barotropic for the same wave length.
- 幅:1326px
- 高さ:193px
- ファイルサイズ:10.9KB
- 幅:1307px
- 高さ:194px
- ファイルサイズ:10.9KB
- 幅:1346px
- 高さ:193px
- ファイルサイズ:10.4KB
- 幅:1080px
- 高さ:600px
- ファイルサイズ:26.6KB
- 幅:1367px
- 高さ:140px
- ファイルサイズ:7.4KB
- 幅:1281px
- 高さ:124px
- ファイルサイズ:8KB
- 幅:1326px
- 高さ:113px
- ファイルサイズ:6KB
- 幅:1325px
- 高さ:315px
- ファイルサイズ:15.4KB
- 幅:1319px
- 高さ:567px
- ファイルサイズ:33.1KB
- 幅:1347px
- 高さ:180px
- ファイルサイズ:11.4KB
- 幅:1407px
- 高さ:206px
- ファイルサイズ:13.2KB
- 幅:1233px
- 高さ:187px
- ファイルサイズ:13.4KB
- 幅:467px
- 高さ:153px
- ファイルサイズ:4.5KB
- 幅:1440px
- 高さ:307px
- ファイルサイズ:21.2KB
- 幅:1293px
- 高さ:204px
- ファイルサイズ:14.1KB
- 幅:1733px
- 高さ:533px
- ファイルサイズ:39.8KB
3. Elementary considerations
If we confine our discussion to a free wave which is periodic with respect to time;
[Fig.107_138_05] is reduced to
式(14) [Fig.107_138_06]
and the volume flux is given by
式(15) [Fig.107_138_07]
and
式(16) [Fig.107_138_08]
where
式(17) [Fig.107_138_09]
The boundary condition (10) is written as
式(18) [Fig.107_138_10]
From (15) and (16), it is clearly noticed that the components of the volume flux are related not only to the surface slope in the corresponding direction but also to the transverse slope because of the effect of the rotation, and (18) shows that the wave number in the x-direction is related to the wave number in the y-direction in contrast to the case without rotation in which the wave numbers in x and y directions are independent of each other.
An elemetary solution of (14) is given by
式(19) [Fig.107_138_11]
where A is an arbitrary constant and Sm^{2} = k^{2}-m^{2}. If we take m real (periodic waves in the y-direction) the character of the wave changes from the Kelvin to the Poincare type according to the condition m^{2} <= k^{2} or m^{2} > k^{2}.
When the period of a wave is larger than half a pendulum day, plane harmonic waves of Poincare type with uniform conditions along the wave crests cannot exist because k^{2} becomes always negative.
Now consider a simple example of reflection when a train of plane harmonic waves (k^{2}>= m^{2}) is incident upon a straight boundary.
The incident wave may be given by
式(20) [Fig.107_138_12]
and the reflected wave which satisfies (18) is given by
式(21) [Fig.107_139_01]
where
式(22) [Fig.107_139_02]
The reflected wave train shifts phase with respect to the incident wave by \phi\ but the amplitude renlaiIls the same. If we introduce the incident angle \theta\ by measuring the incident angle clockwise from the seaward normal to the coast-line, the phase angle \phi\ is expressed by
式(23) [Fig.107_139_03]
Thus, according to the sign of \theta\, the phase of the reflected wave either leads or lags with respect to the case of no rotation.
At \theta\=0, which is the case of normal incidence,\phi\=0 and no phase change occurs. For [Fig.107_139_04],we have \phi\=+-\pi\ and the incident wave \zeta^{in}\ and the reflected wave \zeta^{ref}\ cancel each other completely as is already remarked by CREASE (1958).
- 幅:374px
- 高さ:133px
- ファイルサイズ:3.4KB
- 幅:1299px
- 高さ:134px
- ファイルサイズ:7.8KB
- 幅:1320px
- 高さ:219px
- ファイルサイズ:12.7KB
- 幅:1333px
- 高さ:234px
- ファイルサイズ:13KB
- 幅:1326px
- 高さ:166px
- ファイルサイズ:8.6KB
- 幅:1299px
- 高さ:207px
- ファイルサイズ:13.4KB
- 幅:1340px
- 高さ:124px
- ファイルサイズ:9.3KB
- 幅:1334px
- 高さ:130px
- ファイルサイズ:10.3KB
- 幅:1327px
- 高さ:140px
- ファイルサイズ:9.5KB
- 幅:1333px
- 高さ:213px
- ファイルサイズ:11.7KB
- 幅:1333px
- 高さ:200px
- ファイルサイズ:10.1KB
- 幅:359px
- 高さ:179px
- ファイルサイズ:4KB
4. Reflection of a train of circular waves which is periodic in time
To represent divergent circular waves originated at a point source (x0,y0), we take the free-space Green’s function of (14) as follows:
式(24) [Fig.107_139_05]
where H0, is the zero order Hankel function of the first kind (MORSE & FESHBACH,1953,P.891).
An integral representation of (24) is given by
式(25) [Fig.107_139_06]
To satisfy the boundary condition (18) at x=0, we take a source at the image point with respect to the boundary. Here we should take the phase change of the reflected wave (22) into consideration to account for the effect of the rotation. Thus,we have
式(26) [Fig.107_139_07]
where
[Fig.107_139_08]
Exact integration of (26) seems to be rather complicated but an approximate form of the integral may be derived for large values of kr with [Fig.107_139_09] by means of the method of steepest descents (See Appendix I).
The result is
式(27) [Fig.107_139_10]
where
[Fig.107_139_11]
and
式(28) [Fig.107_139_12]
The first term in the right hand side of (27) may be formally written as
[Fig.107_139_13]
and,denoting \tehta\ the incident angle of circular waves at (0,y), along the boundary, we have
[Fig.107_139_14]
This relation of the phase difference between the incident and reflected waves is exactly the same as that given for plane waves (23).
Therefore,the incident and reflected waves tend to chancel each other along the boundary for a large value of kr since \theta\ approaches +- \pi\/2.
The term G1^{k}, which comes from the contribution of the pole in (26) and represents the Kelvin waves generated by the complicated reflection of circular waves at the straight boundary, remains without attnuation for y < 0 but for y > 0 no such waves exist. This Kelvin wave depends on x0, and the ratio f/\omega\ so that the amplitude is negligible for a large value of x0 and/or for a small value of f/\omega\.
- 幅:1387px
- 高さ:313px
- ファイルサイズ:17.3KB
- 幅:1461px
- 高さ:595px
- ファイルサイズ:27.6KB
- 幅:1553px
- 高さ:639px
- ファイルサイズ:33.7KB
- 幅:993px
- 高さ:227px
- ファイルサイズ:16KB
- 幅:867px
- 高さ:140px
- ファイルサイズ:8.3KB
- 幅:1600px
- 高さ:487px
- ファイルサイズ:32.5KB
- 幅:1727px
- 高さ:480px
- ファイルサイズ:32.4KB
- 幅:673px
- 高さ:133px
- ファイルサイズ:6.1KB
- 幅:686px
- 高さ:213px
- ファイルサイズ:6.6KB
- 幅:1200px
- 高さ:213px
- ファイルサイズ:10.5KB
The boundary wave of Klelvin type generated by an atmospheric disturbance
To investigate the unattenuated boundary wave which is supposed to be generated by atmospheric disturbances, the contribution of Kelvin wave G^{k} for an instaneous point source is evaluated.
Since we know the Kelvin wave caused by an periodic point source, G^{k} may be expressed by
式(29) [Fig.107_140_01]
From the condition (28) on G1^{k}, the limits of integration with respect to \omega\ should be -\alpha\ and \alpha\ where
[Fig.107_140_02]
However, we may put \alpha\ →\infty\ for a large distance in the y-direction from the source, namely, for(x+x0)/Iy-y0I << 1.
Thus, the integral (29) can be evaluated* to be
[Fig.107_140_03]
for
[Fig.107_140_04]
and
式(30) [Fig.107_140_05]
for
[Fig.107_140_06]
With the aid of (30), the contribution of boundary waves of kelvin type in (13) is given by (See Appendix II)
式(31) [Fig.107_140_07]
Now, we can take, without loss of generality, the region of the atmospheric disturbance to be completely enclosed by a rectangle formed by points(0,0),(b,0)(b,a),and (0,a). Thus, Fx and Fy are always zero on and outside of the lines formed by these points. Only limitation imposed on the region of atmospheric disturbances is (x+b)/lyl << 1 as a consequence of the condition assumed to drive G^{k}. After some manipulations (Appendix III),(31) can be transformed to give
式(32) [Fig.107_140_08]
Substituting (30) for G^{k}, we have finally,
式(33) [Fig.107_140_09]
for y<0
where
式(34) [Fig.107_140_10]
Here T stands for a kind of the weighted average in the x-direction of the y-component of the external force and the weight is maximum at the coastal boundary and decreases exponentially with distance from the boundary.
It is remarkable that only the y-component (parallel to the boundary) contributes to the generation of the Kelvin wave and no effect of the on-and off-shore component of the external force appears in the final result.
For an arbitrary distribution of external forces,(33) can be integrated rather easily by a graphical method. However, to show the situation more clearly by a simple model, we assume the distribution of the wind stress as follows:
式(35) [Fig.107_141_01]
for 0 < t0 <= tm, and for 0 < x0 < b, 0 < y0 < a and Fy=0 for the outside of the above region.
Then the Kelvin wave on the boundary for a large distance to the right from the source region is given by*
式(36) [Fig.107_141_02]
Therefore, along the boundary, we have the following solutions for a < ctm:
式(37) [Fig.107_141_03]
This solution is graphically shown in Fig.2.
Fora > ctm,the intervals of t will be;
[Fig.107_141_04]
and [Fig.107_141_05]
The amplitude for (3) will be
式(38) [Fig.107_141_06]
and the remaining part will be the same expressions to the case of a < ctm.
It may be noticed that for a < ctm the maximum elevation of the water surface is given by
[Fig.107_141_07]
and this formula resembles to a wind set up formula in the theory of storm surges where the winds perpendicular to the coastline contribute to the elvation in contrast to the present case where the winds parallel to the coast is important.
* The line of integration is taken to be from -\infty\ +i\epsilon\ with Small positive quantity \epsilon\,to be consistent with the integral representation of Dirac’s delta function.
- 幅:1407px
- 高さ:400px
- ファイルサイズ:21.4KB
- 幅:853px
- 高さ:227px
- ファイルサイズ:11.1KB
- 幅:1141px
- 高さ:393px
- ファイルサイズ:20.6KB
- 幅:1014px
- 高さ:126px
- ファイルサイズ:8.4KB
- 幅:1354px
- 高さ:106px
- ファイルサイズ:7.6KB
- 幅:1181px
- 高さ:107px
- ファイルサイズ:8.4KB
- 幅:1741px
- 高さ:787px
- ファイルサイズ:44.2KB
- 幅:1414px
- 高さ:420px
- ファイルサイズ:20.1KB
- 幅:1354px
- 高さ:414px
- ファイルサイズ:17.3KB
- 幅:1327px
- 高さ:321px
- ファイルサイズ:13.6KB
- 幅:1339px
- 高さ:133px
- ファイルサイズ:8KB
- 幅:1559px
- 高さ:300px
- ファイルサイズ:17.3KB
- 幅:1586px
- 高さ:1400px
- ファイルサイズ:72KB
- 幅:1665px
- 高さ:195px
- ファイルサイズ:14.5KB
- 幅:1093px
- 高さ:214px
- ファイルサイズ:11.4KB
- 幅:1353px
- 高さ:187px
- ファイルサイズ:8.2KB
- 幅:1119px
- 高さ:174px
- ファイルサイズ:9.4KB
- 幅:1286px
- 高さ:1107px
- ファイルサイズ:43.9KB
6. Interpretation in terms of a model of two-layer ocean
It may be of interest to interprete the result in the previous section in terms of a two-layer model of the ocean.
If we put R1 and R2 as the solutions of (7) for the representative elevation of the water level in the barotropic and baroclinic modes respectively, the surface elevation h may be expressed by (KAJIURA,1958)
[Fig.107_141_09]
where (\epsilon\=\rho\1’\rho\2 is the density ratio of water in the upper and lower layers. D1 and D2 are the thickness of the upper and lower layers respectively with the total depth D:D=D1+D2.
The minimum phase velocities of barotropic and of baroclinic waves are respectively
[Fig.107_142_01]
For the example in the previous section, the contribution h1 of the barotropic mode for the surface elevation h is
[Fig.107_142_02]
On the other hand, the contribution h2 0f the baroclinic mode for the surface elevation is
[Fig.107_142_03]
The density ratio \epsilon\ does not enter into the expression explicitly. The ratio of the contributions from the barotropic and the baroclinic modes is given by D1/D2. This shows that if D1/D2 is very small, the amplitude of the Kelvin wave of the barotropic mode is negligible compared with that of the baroclinic (internal) mode, although, these two modes of Kelvin waves have different velocities of propagation and will be completely separated for large values of t.
Take realistic values for the wind field as follows
I\tau\yI=5 c.g.s. (wind speed of about
13m/sec)
a=200km
D1=100m
D2=4000m
The surface elevation will be about 10cm and the deformation of the internal boundary will be about 40m. If we assume (1-\epsilon\)=2x10^{-3} the velocity of the internal Kelvin wave for the present model is c2〜1.4m/sec or about 100km/day. Thus, for the atmospheric disturbance with the duration of about 2days, the duration of the water level variation from the beginning to the end will be about 4 days.
* Because of the sharp decrease of the weight function on \tau_{y}\,b may be replaced by \infty\ without serious error.
- 幅:867px
- 高さ:207px
- ファイルサイズ:9.3KB
- 幅:1214px
- 高さ:200px
- ファイルサイズ:13KB
- 幅:707px
- 高さ:200px
- ファイルサイズ:7.5KB
- 幅:747px
- 高さ:394px
- ファイルサイズ:17KB
7. Summary and concluding remarks
It is shown in§4 that the circular waves radiating from the point source are reflected at the straight boundary through complicated processes and generate boundary waves of Kelvin type at a large distance to the right (facing offshore in the northern hemisphere), the amplitude of which depends on f/\omega\ and the distance of the source from the boundary.
In§5 the boundary wave of Kelvin type generated by an atmospheric disturbance is discussed, and it is found that the external force in the y-direction, say, the wind stress parallel to the boundary, is important and the force in the x-direction, namely, on-or offshore winds do not contribute to the generation of Kelvin waves. Morever, the contribution of the parallel winds in the region close to the boundary is most effective as is manifested by the weight function exp (-fx0/c).
Therefore, the wind field of the distant region from the boundary does not play a role in the Kelvin wave generation, although the approximation employed does not permit the rigorous discussion if (IyI+y0)/(x十x0o) is not so large. In §6, an application of the result to a two-layer model of the ocean is attempted.
In the model employed, the Kelvin wave of the barotropic mode is small compared with that of the baroclinic mode if the ratio of the thickness of the upper layer to that of the lower layer is small. The order of amplitude of the internal Kelvin wave generated by an ordinary storm is about 10cm at the surface and the wave is of the solitary type with the duration of a few days.
In the application of the theory to the actual ocean, many factors, which are not adequately taken into account in the present paper, should be considered. For example, the structure of the water body in an real ocean may be very far from the two-layer model and in many places the existence of a continental shelf and a steep slope connecting the shelf to the deep ocean proper complicates the generation and propagation of boundary waves of various kinds. Actually, on the continental shelf, the boundary waves of Kelvin type might be only a minor part of the spectrum of varieties of boundary waves such as edge waves.
However, it may be of some interest to mention that an analysis of day to day variations of sea level along the Japanese Islands by SHOJI (1961) seems to indicate that the maximum or minimum of the daily mean sea levels appears to propagate sometimes to the west along the Pacific coast and in these occasions the variations of the mean sea level in the original area are caused by the wind field parallel and close to the coastline. This finding by SHOJI is in good qualitative agreement with the present theory.
References
CREASE, J.(1956):Long waves on a rotating earth in the presence of a semi-infinite barrier. J.
FLUID MECH.,1,86-96.
CREASE,J.(1958):The propagation of long waves into a semi-infinite channel in a rotating system
J.FLUID MECH.,4,306-320.
ICHIYE, T.(1958): The response of a stratified, bounded ocean to variable wind stresses. Ocn.
Mag.,10,19-63.
KAJIURA, K.(1958): Response of a boundless two-layer ocean to atmospheric disturbances. Dis-sertation, A&M College of Texas,139 pp.
LAMB, H.(1932):Hydrodynamics,6th Ed., Cam-bridge University Press, Cambridge.
MORSE, P. M. ane H. FESHBACH (1953): Methods of Theoretical Physics, Vol.1, McGraw-Hill Book
Co., New York.
SHOJI, D.(1961): On the variation of the daily mean sea levels along the Japanese Islands.
Journ. Ocn. Soc. of Japan,17 141-152.
TAYLOR, G.I.(1922): Tidal oscillations in gulfs and rectangular basins. Proc. London Math.
Soc.,2nd Ser.20,148-181.
VERONIS, G. and H. STOMMEL(1956): The action of variable wind stresses on a stratified ocean.
J.Mar.Research,15,45-75.
VERONIS, G.(1956): Partition of energy between geostrophic and non-geostrophic oceanic motions.
Deep Sea Research,3,157-177.
YOSHIDA, K.(1960):The oceanic waves of days to month’s periods. Rec. Oceanogr. Works in Japan,5,11-24.
Appendix I
The integral defined by
式(I-1) [Fig.107_143_01]
may be transformed into,
[Fig.107_143_02]
where
[Fig.107_143_03]
The shaded area in Fig. I-1 is the region of convergence with a pole located at tan\theta\ = if/\omega\ and the line of integral \Gamma\ is assumed to avoid the pole to the right as shown in the figure. In the deformation of the contour from \Gamma\ to W, which crosses the saddle point on the \Theta\-plane and follows the steepest path according to the method of steepest descent (Endelyi;Asymptotic Expansions, Dover Publ.,
Inc.,1956,39-41). W crosses \Gamma\ in the upper half plane when -\pi\ < \delta\ < -\pi\/2 and the pole is crossed whentan^{2}\delta\ < f^{2}/(kc)^{2}. Therefore
the contribution of the pole should be taken into account if (y-y0) < 0 and (y-y0)^{2}/(x+x0)^{2} > (\omega\^{2}-f^{2})/f^{2}.
The integral along the steepest path near the saddle point gives nothing but the asymptotic form of the Hankel function of the first kind with the coefficient evaluted at the saddle point. It is remarked here that even if k is imaginary (f > \omega\), we can arrive at the similar conclusion by means of a slightly different transformation.
- 幅:1446px
- 高さ:573px
- ファイルサイズ:35.8KB
- 幅:1512px
- 高さ:307px
- ファイルサイズ:23.2KB
- 幅:1305px
- 高さ:320px
- ファイルサイズ:21.7KB
- 幅:1199px
- 高さ:1399px
- ファイルサイズ:71.6KB
Appendix II
From (13) the integral in question is originally written as
式(II-1) [Fig.107_144_01]
The substitution of the boundary condition (10) yields
式(II-2) [Fig.107_144_02]
where
式(II-3) [Fig.107_144_03]
Now, by meanshave ofthe partial integration,we have
式(II-4) [Fig.107_144_04]
where \alpha\ and \beta\ are to be determined by the condition for G so that on x0=0 the region of integral is enclosed by a curve defined by
式(II-5) [Fig.107_144_05]
and
t0=0
as shown Fig. II-1.
The first integral in the right hand side of (II-4) may be written, in terms of a line element on the boundary as follows:
式(II-6) [Fig.107_144_06]
where ds is taken positive in the clockwise direction and l is the total length on the boundary from (\alpha\0,0) to (\beta\0,0).The second integral is identically zero because of the boundary condition and the third integral becomes, after exchanging the order of integration, and taking the initial condition into account,
式(II-7) [Fig.107_144_07]
Therefore (II-4) is reduced to
式(II-8) [Fig.107_144_08]
It can be easily shown from (30) that
式(II-9) [Fig.107_144_09]
since x << ly-y0I and c(t-t0) = -(y-y0). Therefore we can conclude that the contribution of G^{k} on I is
式(II-10) [Fig.107_145_01]
Strictly speaking, in (II-6) and (II-7) we have divergent integrals sillce G and \delta\G/\delta\x0 include divergent terms on \Gamma\. However, we may apply the technique developed by Hadamard on the finite part of an infinite simple integral (Hadamard;lectures on Cauchy’s Problem, Dover Publ., Inc.,1952, p.133).
- 幅:1573px
- 高さ:420px
- ファイルサイズ:25.8KB
- 幅:1586px
- 高さ:313px
- ファイルサイズ:17.1KB
- 幅:1607px
- 高さ:433px
- ファイルサイズ:25.3KB
- 幅:1680px
- 高さ:627px
- ファイルサイズ:40.8KB
- 幅:1334px
- 高さ:114px
- ファイルサイズ:8.8KB
- 幅:1547px
- 高さ:726px
- ファイルサイズ:38.7KB
- 幅:1607px
- 高さ:713px
- ファイルサイズ:43.3KB
- 幅:1520px
- 高さ:620px
- ファイルサイズ:32.5KB
- 幅:1384px
- 高さ:315px
- ファイルサイズ:17.4KB
- 幅:1254px
- 高さ:1033px
- ファイルサイズ:44.5KB
- 幅:1040px
- 高さ:299px
- ファイルサイズ:13.2KB
Appendix III
The fi rst term on the right hand side of (31) is trahsformed into
式(III-1) [Fig.107_145_02]
where d’= c(t-t0,)+y, y < 0.
Taking the condition that Fx=Fy=0 along , x0=b and y0=0,(31) may be reduced to
式(III-2) [Fig.107_145_03,04]
The boundary values at x0=0 are cancelled with each other.
The last integral in (III-2) is tranformed, by making use of the explict representation of G^{k},into
[Fig.107_145_05]
and the exchange of the order of integration yields
[Fig.107_145_06]
Since Fx=Fy=0 at t0=0 and
[Fig.107_145_07]
we have finally
式(III-3) [Fig.107_145_08]
Here, it can be shown that the term in the bracket in the right hand side of (III-3) vanishes. Furthermore, the first term can be transformed by exchanging the order of integration to dive finally,
式(III-4) [Fig.107_145_09]
- 幅:1593px
- 高さ:1679px
- ファイルサイズ:99.6KB
- 幅:1566px
- 高さ:533px
- ファイルサイズ:26KB
- 幅:1519px
- 高さ:493px
- ファイルサイズ:31.1KB
- 幅:1606px
- 高さ:294px
- ファイルサイズ:21.5KB
- 幅:1673px
- 高さ:267px
- ファイルサイズ:20.6KB
- 幅:947px
- 高さ:213px
- ファイルサイズ:10.7KB
- 幅:1506px
- 高さ:1006px
- ファイルサイズ:55.2KB
- 幅:1453px
- 高さ:426px
- ファイルサイズ:20.4KB
27.湾水振動におよぼす防波堤の効果 地震研究所梶浦欣二郎 (昭和37年11月27日発表一昭和38年3月29日受理)
§1.はしがき
最近国土開発あるいは津波,高潮の災害を防ぐ目的で方々の湾において大防波堤の建設が計画され,防波堤建設によつておこる湾内潮汐の変化あるいは津波,高潮の波高変化を知るための模型実験が行なわれている.1).2).3) 又一方では,東京湾共の他における高潮についての数値実験が電子計算機を用いて行なわれ,その際には大防波堤建設によつて防波堤内部の高潮の波高がどの程度軽減されるかも計算された.4).5) ところで,模型実験あるいは数値実験で得られたものは個々の湾に対する結果であるが,これらの結果のうちどの程度までが一般的な結論であるかは明らかでない.このことは,例えば模型実験の結果が,その問題に最も適当と思われるパラメータについて表示されていないところにもあらわれている.
今ここでのべようとするのは,Love (1959)6)の行つたと同様に,極めて簡単な湾のモデルを考え,防波堤の開口部では水理学で普通に用いられる水位流量の関係式を使用して理論的に潮汐,津波,及び高潮に対する防波堤の効果を計算し,その結果がどの程度定量的に今迄に行なわれた模型実験の結果を説明するか,また逆にどのような点で,今までの模型実験結果を一般的なものとみることが出来るかということである.勿論,模型実験の目的はこの様な一般的結論よりは個々の湾についての結果そのものであるが,どの様な点にその湾の特殊性があるかを知るには,簡単な理論的推察と実験結果とを対応させ,その違つている点にもつと重点的な考察を進めるべきであると思われる.
また一方では,この様な簡単なモデルによる考察は,複雑な地形をした湾内の海水振動の数値実験を行なう場合に,どの程度の項まで含めた基礎方程式を用いるのが適当であるかの目安をも与えることが出来る.これに関係して線型方程式を基礎とした理論から導かれる港湾副振動についての性質を現実に適用することには限界があることが判る.即ち,防波堆の開口部が小さくなるとそこでの非線型の効果が最も重要となるために,Miles and Munk (1961)7) ののべた様な,防波堤の開口部を狭くすることは副振動を小さくすることにはならないという港湾副振動についてのパラドックスは防波堤開口部の面積が可成り大きくないと適用しがたい.
1) 長崎海洋気象台「模型実験による有明海の潮せきに関する研究」第1報,研究時報12 (1960),261.第2報,研究時報13 (1961),83.
2) 岩崎敏夫・三浦晃「気仙沼湾の津波の模型実験について」第8回海岸工学講演会講演集 (1961),51.
3) 樋口明生・吉田幸三「名古屋港の水位変動の特性について」第9回海岸工学講演会講演集(1962),34.
4) 宮崎正衛「東京湾高潮に対する防波堤効果の計算について」産業計画会議資料第217号 (1961),171.
5) 宇野木早苗・磯崎一郎「防潮堤開口部の流速計算についての一考察」第9回海岸工学講演会講演集 (1962),7.
6) R.W. LOVE,「Tidal response of a bay with a constricted opening to the sea.」Thesis for M. S., A and M College of Texas, Dept. of Oceanogr.,(1959),66 p.
7) J.MILES and W. MUNK,「Harbor Paradox,」Proc., ASCE,87, No. WW 3,(1961), 111.B. L. MEHAUTE and B. W. WILSON,「Discussion,」Proc., ASCE,88, No. WW 2, (1962),173.
§2.基礎方程式
今,湾水の運動が一次元的なモデルで近似出来るとし,長波の近似が許され,防波堤の開口部を除いては線型近似が出来また海底摩擦や渦によるエネルギー消耗はないと仮定すると,運動方程式及び連続の式は次の様に書ける.
式(1) [Fig.107_147_01]
式(2) [Fig.107_147_02]
ここで,Q:流量,\eta\:平水面からの水位上昇量,x:水平座標, t:時間,g:重力加速度, h:水深,b:水路巾,.F:湾内水に作用する外力,であり,考えている湾のモデルは第1図に示すように,一様な短形断面をもつ水路が防波堤位置において連結されている.座標原点 x=0 は防波堤位置にとり,湾と外海との境界を x=-x1,湾奥の位置を x=x2とし,必要のある場合には外側の湾に対する量には添字1を,内側の湾に対する量には添字2をつけて区別する.今海水の運動による摩擦力は考えていないので,潮汐あるいは津波の問題を取り扱うときにはF=0とするが,高潮を考える場合にはF=\tau_{s}\/\rho\ (\tau_{s}\:風の海面応力,\rho\:水の密度)を採用する.
防波堤開口部 (x=0) の条件としては,水理学で一般に使用される水位差と流量との関係式を利用する.即ち,
式(3) [Fig.107_147_03]
ここで,A0:防波堤開口部面積, k:流量係数であり,複号は (\eta\1-\eta\2) の符号に順ずる.
流量係数kの値は,開口部の形によつてかなり変化することが知られているが,厚みのない板のような薄い防波堤の場合にはk〜0.7程度のようである.8) この式 (3) 自体は,もともと定常流について求められたものであつて今考えるような時間的に変化する流れ,あるいは外力の作用下の流れにも適用出来るかどうかには考慮の余地があるが,完全な運動方程式に立ち帰つて各項の大いさを吟味すれば判るように,周期の長い波動の場合について,開口部の面積が水路の断面積にくらべてかなり小さいとき,また外力については防波堤開口部の奥行きが短いときには成立するとみられる.
x=0 における条件が(3)で与えられると,後は問題に応じて適当な初期条件や境界条件を与えることによつて,外側の湾及び内側の湾に対する水の運動はよく知られた方法で解くことが出来る.今解法の点からみて二つの異つた場合,即ち,外海から伝わる擾乱が完全に周期的な波の場合と弧立波的な波の場合とを分けて考えることにする.高潮の場合のように周期的でない外力が湾水全体に作用するときは後者の方法が便利である.
8) VEN TE CHOW, Open-channel hydraulics (McGraw-Hill Book Co. Inc.,1959), Chap.17
- 幅:1906px
- 高さ:240px
- ファイルサイズ:14.8KB
- 幅:1846px
- 高さ:193px
- ファイルサイズ:12.7KB
- 幅:2206px
- 高さ:166px
- ファイルサイズ:14.6KB
- 幅:1246px
- 高さ:646px
- ファイルサイズ:29.1KB
§3. 湾ロで周期的な水位変動が与えられる場合の湾水振動一潮汐振動
上述のモデルにおいて,湾口(x=-x1)の水位変動が周期的なものとして与えられる場合を考え,F=0とする.このとき境界条件は,
式(4) [Fig.107_148_01]
また湾奥での条件は,
式(5) [Fig.107_148_02]
であり,H:湾口水位変化の振巾,\omega\=2\pi\/T,T:外力の周期である.(5)は湾奥での完全反射の条件であるが,もし湾奥で完全反射が起こらないときにはこの式を変更する必要がある.今のところ,実際の湾奥条件をどのよらにとればよいかの問題は未解決といえるが簡単には,流量が水位勾配に比例するような条件式を使うのが普通である.しかしここでは完全反射の式を採用しておく.
ここで与えた問題の最も簡単な場合として,湾口条件(4)を防波堤直前で与えたときの内湾の海水振動についてはすでにLoveによつて議論がなされている.今のべる理論では,それをやや拡げて外湾の効果をも含めてある.
今(1),(2)式を(3),(4),(5)の条件のもとに解くと(3)が非線型効果を含んでいるためにやや複雑な解が得られる.しかし非線型の作用によつてあらわれる高調波はその振巾が小さいので省略することにすると基本周期のもののみからなる解となる.具体的には(3)のうちに含まれる
式(6) [Fig.107_148_03]
のような展開において右辺第1項のみを残すこと相当する.
結局(1),(2)の解は(3),(4),(5)の条件と(6)の近似のもとに次のように与えられる.
式(7) [Fig.107_148_04]
式(8) [Fig.107_148_05]
ここでH1,\epsilon\1は決定すべき常数, [Fig.107_149_01]である.今求めようとする量は湾奥の振巾及び位相のおくれ \epsilon\2 であつてそれらは次の関係式によつて与えられる.
式(9) [Fig.107_149_02]
式(10) [Fig.107_149_03]
ここで,
式(11) [Fig.107_149_04]
式(12) [Fig.107_149_05]
式(13) [Fig.107_149_06]
式(14) [Fig.107_149_07]
S1:外湾の表面積、 S2:内湾の表面積
(10)から,位相 \epsilon\2 については 0 =< \epsilon\2 =< \pi\12であり湾内振動の位相は常に湾口に比して遅れることが判り,その量はパラメータBによつてきまる.(cosk1x1 < 0 のときは湾口に対する位相のおくれは ( \pi\+\epsilon\2) である.) (11)において,Bの大勢はB0によつて決り,(12)を考慮すると,Bは湾口での波動振巾に比例し,内湾の表面漬の2乗に比倒し,周期及び防波堤開口部面債の2乗に逆比例することが判る.このような傾向は,模型実験の結論ともよく一致し,防波堤開口部の条件が口の狭い場合には本質的に非線型であることを示している.湾の長さと深さとに関係した量k1x1,,k2x2もパラメータBに含まれ,特に(13)をみれば \beta\ が防波堤位置に関係した量であることは明瞭であり,もし防波堤位置を湾口と考えると \beta\=1となる.
また(12)に含まれるH0は(14)で示されるように,防波堤に開口部がないと考えたとき,即ち外湾のみがあるときに,防波堤位置(外湾の湾奥)で期待される水位変化の振巾であつて,B0中には外湾に関する量はこれ以外には入つていない.即ちLoveの計算はH0を与え, \beta\=1とおいた場合に縮当する.
(10)を用いて,パラメータBに対する \epsilon\2(位相)及びcos\epsilon\2(振巾)の変化を計算したものが第2図である.図からも明らかなように,パラメータBの増加とともに内湾の位相は90°まで遅れ,振巾は0まで減少する.
- 幅:1986px
- 高さ:166px
- ファイルサイズ:11.1KB
- 幅:1832px
- 高さ:146px
- ファイルサイズ:8.9KB
- 幅:2380px
- 高さ:200px
- ファイルサイズ:19.7KB
- 幅:2560px
- 高さ:140px
- ファイルサイズ:18KB
- 幅:2106px
- 高さ:127px
- ファイルサイズ:12.1KB
- 幅:766px
- 高さ:147px
- ファイルサイズ:7.2KB
- 幅:2199px
- 高さ:140px
- ファイルサイズ:13.5KB
- 幅:2046px
- 高さ:120px
- ファイルサイズ:10.8KB
- 幅:1760px
- 高さ:140px
- ファイルサイズ:7.4KB
- 幅:2192px
- 高さ:213px
- ファイルサイズ:21.1KB
- 幅:2273px
- 高さ:226px
- ファイルサイズ:21.1KB
- 幅:1880px
- 高さ:107px
- ファイルサイズ:8.1KB
§4. 周期的でない外力の作用による湾水の振動一津波或いは高潮
一次元波動の間題において,過渡的な現象を議論するのに便利な方法の一つとして特性曲線法がある.これは非線型の波動を取り扱うのに広く用いられているが,線型の波動の場合にも他の方法にくらべてきわめて有効に用いることが出来る.この方法では(1)(2)を次のように書き換える.
まず,
式(15) [Fig.107_150_02]
式(16) [Fig.107_150_03]
とおき,水路の巾及び水深が一定であることを考慮すると,
式(17) [Fig.107_150_04]
式(18) [Fig.107_150_05]
となる.ここで [Fig.107_150_06] は長波の速度である.
今初期条件として t=0における \eta\ 及びQ (即ちZ+及びZ-) をxの函数として与える.境界条件としては x=0で(3)を,x=x2では(5)を与える. x=-x1の条件としては例えば外海からの進行波を問題にするときは,§2のモデルにおいて外湾は無限に長いと考え, x1\=infty\ とし,x=-x1 では波の反射がないとするのも一方法である.また湾内の高潮のみを問題にするときには,湾口x=-x1での条件を例えば \eta\1=0とする.第3図は津波が内湾に向つて進んで来る場合の一例を示す.
ここで,防波堤開口部即ちx=0における条件が,Z+及びZ-を用いてどのように書けるかが問題になるが,そのまえにまず防波堤のない場合を考える.このときx=0で水路の断面積が突然変るために波の部分反射が予想される.これを表現すると,
式(19) [Fig.107_150_07]
であるから,\alpha\=(b2c2)/(b1c1)として,
式(20) [Fig.107_151_02]
式(21) [Fig.107_151_03]
となる.ここでxの正方向(湾奥に向う)に進む波については 2/(1十\alpha\)が透過係数,(1-\alpha\)/(1+\alpha\)が反射係数を与えている.
これに反し,防波堤のある場合には非線型の条件が入つているためにことがらが面倒になる,即ち反射あるいは透過の割合が波の振巾によつて異なるのである.この場合には,Zの代りに,
[Fig.107_151_04]
h1:外側の水路の水深,
A1:外側の水路の断面積
で定義される無次元量Yを使つて,
式(22) [Fig.107_151_05]
式(23) [Fig.107_151_06]
と書ける,ここで複号は(Y1+ + Y2-)の符号に順ずる.したがつて,水位 \eta\ 及び流量Qは
式(24) [Fig.107_151_07]
となる.また(17),(18)式は
式(25) [Fig.107_151_08]
式(26) [Fig.107_151_09]
と変形される.
以上で特性曲線法による津波,高潮計算を行なう際の防波堤開口部の条件は明らかになつた.ここで注意すべきことは,長さの次元をもつパラメータが [Fig.107_151_10]という型で入つていろことであつて,前述の周期的な振動の場合のパラメータBとは異なっている.今の場合,例えば初期条件として入射波の波高 \eta\*を与えたとすると,ある一定の \eta\*/Kの値に対してなされた計算はそのままでは一般に他の条件の場合に拡張出来ないが結果の解訳にあたつて \eta\/h1 と [Fig.107_152_01] とがあるきまつた関係を満すかぎりでは異なつた初期波高の場合にも適用される.例えば初期波高が2倍になることは,防波堤の力学的な効果としては初期波高に変化なくA0が1/√2に減少することに相当している.
- 幅:2386px
- 高さ:1799px
- ファイルサイズ:145.4KB
- 幅:1906px
- 高さ:133px
- ファイルサイズ:9.8KB
- 幅:1886px
- 高さ:140px
- ファイルサイズ:9.3KB
- 幅:1966px
- 高さ:247px
- ファイルサイズ:15.5KB
- 幅:1966px
- 高さ:220px
- ファイルサイズ:14.3KB
- 幅:367px
- 高さ:127px
- ファイルサイズ:3.9KB
- 幅:2073px
- 高さ:147px
- ファイルサイズ:11.1KB
- 幅:1593px
- 高さ:1473px
- ファイルサイズ:75.3KB
- 幅:1280px
- 高さ:213px
- ファイルサイズ:13.5KB
- 幅:1247px
- 高さ:227px
- ファイルサイズ:14.5KB
- 幅:1280px
- 高さ:266px
- ファイルサイズ:18.2KB
- 幅:2434px
- 高さ:240px
- ファイルサイズ:20.7KB
- 幅:2200px
- 高さ:219px
- ファイルサイズ:18.5KB
- 幅:2227px
- 高さ:260px
- ファイルサイズ:19.6KB
- 幅:1980px
- 高さ:253px
- ファイルサイズ:15.9KB
- 幅:2014px
- 高さ:227px
- ファイルサイズ:15.9KB
- 幅:867px
- 高さ:193px
- ファイルサイズ:13.1KB
- 幅:560px
- 高さ:193px
- ファイルサイズ:8.4KB
§5.有明海の潮汐模型実験結果との比較
有明海は潮汐振動の大きいところとして日本では有名であるが,ここに大防波堤をつくつた場合に内湾の潮汐がどのように変化するかを調べる目的で長崎海洋気象台が中心となつて模型実験が行なわれた.1) このとき行なわれた実験は,第4図に示すようなI,II,III,IV,Vの位置に防波堤を建設する場合,防波堤の開口部を順次に狭めて行つた際に有明海沿岸の観測点で潮汐振巾及び位相がどのように変化するかを調べたものである.(このとき竹崎の南側にある諫早湾は図に示した線で締切られている.)この結果に対して上述の理論的考察をあてはめるために,ここでは有明湾口に位置する口の津の潮汐振巾及び位相を基準にとつて湾奥、部の竹崎,住の江の潮汐振巾及び位相の変化を実験結果からとり上げる.また,防波堤開口部の面積が湾の断面積に比してかなり小さくないと今の理論は適用出来ない筈のものであるから,開口部面積が約5×10^{4}m^{2} (全断面積に対する開口部面積の比は約0.1〜0.2)以下のところを議論する.開口部面積がかなり大きいときには,条件(3)のような水位流量関係から考えられる非線型の効果だけでなく,水位,流量の時間的変化が重要な役割を果していることが実験にもあらわれている.
さて,理論計算にあたつて,まず有明海附近の海図から表1にあげた諸量を計算しておく.(ここで,水深としては,「海図水深」+1.5mを平均水面からの水深と仮定し,その算術平均として,全体の平均水深を計算した.)これらの量及び潮汐周期(12.4時間),湾口振巾(1.5m),流量係数(0.7)を用いると,湾口の漸汐振巾に対する湾奥の潮汐振巾の比,及び位相の遅れは容易に計算出来る.
第5図にはかくして計算された半日潮の潮汐振巾及び位相について,理論と実験との比較が示されている.口の津の漸汐振巾 \eta\0 を基準とした湾奥の住の江,竹崎の漸汐振巾\eta\/\eta\0 を防波堤設置場所I,II,III,Vについてくらべてみると,防波堤断面が5×10^{4}m^{2}以下では理論値と実験値とがかなりいい一致を示していることが判る.これに反して,断面積が10^{5}m^{2}では,その違いが大きく,初めに予想した通り,条件式(3)は不適当であることが判る.断面IIについては,理論値と実験値との差がかなり現われているが,この原因については明らかでない.I及びIIIの断面について,理論値と実験値とが,かなりよく一致しており,またIIとIIIの実験値がほとんど一致することから考えると, IIの断面の実験には何等か特別な条件が加わつているとみられる.IVの断面については正確な実験値が示されていないので議論が出来ない.Vでは住の江の実験値が示されていないので,島原での値を加えてあるが,一般に実験値よりも理論値の方が小さく,防波堤の効果としては大きく現われている.この違いもいろいろの原因によると思われるが,特に流量係数(k)に関係しているかも知れない.
位相については,II及びIIIの位置に防波堤がある場合について,口の津に対する竹崎の位相のおくれのみが示されているが,もともと実験において,位相を正確にきめることが困難であることを考慮に入れると,防波堤開口部の断面積が2.5×10^{4}m^{2}以下ではかなりよく合つているといえさうである.ただ理論では摩擦の効果が無視されているから,竹崎に対する住の江の潮汐位相のおくれのような場所による位相のずれは期待されない.現実には,有明海では摩擦の効果が相当に顕著であつて,潮汐の調和分解を行うと,半日潮の位相が湾奥に向うにつれて順次におくれていることが明瞭にみられる.
理論では一次元の仮定と摩擦の効果を無視したことから考えると,個々の波の形の比較では,今のモデルによる結果と,現実をうつした模型とではかなりの違いが出るはずで,実験値と理論値で1割程度の違いは当然期待されてよい.そのことを考慮に入れると,実験値と理論値とは,防波堤開口部面積が小さいときにはよく一致しているといえそうである.即ち,§3でのべたパラメータ・Bの大小によつて,振巾,位相の変化の大勢を説明出来ることになる.第6図で示したように,有明海の規模の湾では防波堤が湾口近く設置されるほど,同一の開口部面積に対して内湾における潮汐減衰の効果が著じるしく,例えば口の津の半日潮汐振巾1.5mのとき,Vの位置の防波堤を建設して,開口部面積2×10^{4}m^{2}のとき,湾奥の漸汐振巾は30cm程度ということになり,防波堤のない,現在の状態では湾奥の潮汐振巾が約2.5mあることからみると実に1/8程度の振巾に減衰していることになる.このとき口の津の防波堤開口部を通る海水の最大流速は約7ノットに達すると思われる.
- 幅:2060px
- 高さ:1672px
- ファイルサイズ:139KB
- 幅:1813px
- 高さ:799px
- ファイルサイズ:68.7KB
- 幅:2953px
- 高さ:2465px
- ファイルサイズ:271.6KB
- 幅:1293px
- 高さ:1127px
- ファイルサイズ:58.2KB
§6.気仙沼湾の津波模型実験との比較
1960年5月のチリ地震津波によつて東北地方太平洋岸のかなり大きな湾の最奥部で津波被害があり,それを防ぐために各所で湾口防波堤を設置するという案が出されている.
その基礎資料の一つとして,気仙沼湾の湾口に防波堤を設置した場合に,果して津波の波高減衰に対して効果があるかどうかということが模型実験によつて調べられた.2)
模型実験では短形水漕内に気仙沼湾と外海の一部とをうつし,これに単一な押し波又は引き波を進行させ,湾の入口に設けられた防波堤によつて湾内の波高分布がどうなるかが測定されている.第7図には気仙沼湾の地形図及びこれに対応すると仮定された理論的な一次元モデルを示してあり,このモデルについて§4の方法によつて計算を行なう.
まず,湾口の問題を考える.一次元モデルで実際には二次元的な現象を扱かうのであるから,幾分基礎のあやふやになる点のあるのは止むを得ないが,気仙沼湾入口の地形及び一般に波長の長い波の湾口附近での行動から推察して,湾の巾の2倍の巾をもつた水路が外海を代表すると仮定し,ここを伝わる波が湾口で部分反射をおこしながら湾内に入射すると考える.
水深は,地形図からみて湾内で約10m,湾口直前の外海で約25mとなつているから,水深の比を1/2.5とする.即ち断面積の比はA2/A1〜1/5であり波の反射に関する量 \alpha\〜0.3がきまる.湾の長さに対しては,今湾口から湾奥まで長波の進むに要する時間をaとすると,湾水の基本振動周期は4aであり,aをパラメータとして外部から伝わつて来る弧立波の周期(波の初まりから終りまでの時間)を表現することにすると,湾の長さそのものは計算には直接入つて来ない.
まず初めに,防波堤のない場合について,波がどのくらい続いて来襲すると,完全に周期的な波の来る場合と同様な湾内波高が得られるかをみるために,正弦波形の波について計算を行なつたものを第8図に示す.このとき波の周期 na が湾の固有振動周期に近いと振巾が大きくなることは当然期待されるが,今の計算では,海底摩擦がなくても,湾口から外部へ逃げる波のエネルギーが存在するために,平衡の位置には意外に早く達している.
即ち \alpha\=0.3の場合,来襲波が3波程度ですでに周期的な来襲波のときの理論値にほとんど合致していることがみとめられる.この理論値は妹沢(1935)9)の与えたと同様な方法によつて容易に求められる.
今入射波をH1cos(k1x-\omega\t) とすると,湾奥の振巾は
式(27) [Fig.107_156_01]
であり,湾口での振巾は
式(28) [Fig.107_156_02]
で与えられる.ここで [Fig.107_156_03] である.
一山一谷の波の場合には周期が湾の固有振動周期(4a,4a/3,4a/5,…) から離れたところで周期的な波の理論値より一般に大きくなつている.別の言葉でいうと,干渉の作用を受けない方が山が大きい.逆に共鳴周期の近くでは,繰り返しの波の干渉による振巾増大があるわけである.また (28) から判るように,湾口での波の振巾は入射波の振巾H1とは一般に異なつたものであり,湾の大きさや深さ,それに外海とのつながり具合等によつて変化するものであるから,津波の湾口における高さから,外海での津波を推定するときには注意を要する.
次に,湾口に防波堤を設けた場合には来襲弧立波を次のように仮定する.
式(29) [Fig.107_156_04]
ここで [Fig.107_156_05] ,来襲波の周期 T=na である.この初期波はxの正方向に進む進行波と仮定して計算を進める.
§4にものべたように,非線型の効果をもつ防波堤の存在する場合には,計算にあたつて来襲波の波高を与えておく必要がある.今は模型実験の場合に対応させて \eta\0=1.5mをとる.
結果は第9図の通りであるが,ここでは,防波堤のないときの湾奥最大波高 \eta*_{head}\ を基準として防波堤のあるときの湾奥最大波高 \eta_{head}\ をあらわすことにし \eta_{head}\/\eta*_{head}\ がA0/A2 (A0:防波堤開口部面積,A2:内湾の断面積)によつてどう変るかが示されている.入射波の周期が湾の基本振動周期よりも短いとき (n<4) にはこの比 \eta_{head}\/\eta*_{head}\は周期にかかわらず一定であり,n>4 では入射波の周期が長いほど同一の開口部面積に対して防波堤の効果は少ないことがわかる.これからみて明らかなことは,湾水の基本振動周期に比して入射波の周期がきわめて長いときには,防波堤の波高減衰効果は望めないわけで,内水面のそれほど広くない港の港口防波現のようなものは,開口部面積が極端に小さくなければ長周期津波に対してあまり効果のなことは明らかである.
図中には岩崎,三浦による模型実験の結果をも入れてあるが,これらの実験値はすべてn < 4の範囲のものある.計算及び実験の両方において,湾内の波高比 \eta\/\eta\*は場所によつてあまり変らないことが判つているのて,模型実験結果として図に示されているものは,湾内各所で得られた波高比を平均したものである.但し,平均をとる際に,実際の気仙沼湾では湾奥附近に巾のいちじるしく狭くなつたところがあり,摩擦や反射の影響が顕著にあらわれていると思われるので,湾の最奥部数ケ所の実験値は平均から除外してある.
この図をみると,ごく大体の傾向からいつて理論と実験との結果は合つているとみられる.即ち湾口防波堤の開口断面積の減少による湾内波高減少の割合は,今のべている簡単な理論から,かなり定量的にも推察出来る.
一般に津波に対する湾口防波堤の効果という点からみると,例えべ1960年5月のチリ地震津波のような周期1時間に近い波に対しては,湾の基本振動周期が1時間程度ないと,あまり防波堤の効果が発揮されないが,三陸地方でしばしば経験される近地津波を考えると,周期30分程度あるいはそれ以下のことが多いので,同じ開口部面積をもつ防波堤の効果は,がなり小さな湾でもあらわれるであろう.
9) K.SEZAWA,「Growth and decay of seiches in an epicontinental sea,」Bull. Earthq. Res. Inst.,13(1935),476.
- 幅:1367px
- 高さ:1539px
- ファイルサイズ:85.1KB
- 幅:1653px
- 高さ:140px
- ファイルサイズ:14.7KB
- 幅:1440px
- 高さ:140px
- ファイルサイズ:10.2KB
- 幅:786px
- 高さ:120px
- ファイルサイズ:7.5KB
- 幅:2346px
- 高さ:380px
- ファイルサイズ:27.4KB
- 幅:1140px
- 高さ:146px
- ファイルサイズ:11.3KB
- 幅:1080px
- 高さ:1400px
- ファイルサイズ:45.8KB
- 幅:1333px
- 高さ:1126px
- ファイルサイズ:52.3KB
§7.高潮に対する防波堤の効果
§4の方法によつて,高潮の場合に防波堤がどのような効果をもつているかが計算出来るが,その簡単な場合として巾b 及び水深h の一様な短形湾(断面積Aは勿論一定)の内部に防波堤を設置した場合について考える,即ち,第1図のモデルで外湾及び内湾は深さも巾も等しい場合である.今湾の主軸方向に一様な風W (時間的には変化する)が吹く場合に,湾内の高潮が防波堤の設置場所,防波堤の開口部の広さA0,風の吹め始めてから止むまでの時間T等によつてどう変るかを調べる.気圧の効果は今の議論からは一応除外する.
まず,t=0で静止した水面を考え,風の応力は,
式(30) [Fig.107_157_01]
とする.ここで\tau\ は最大風速に対応する応力である.時間は長波の湾口(x=-x1) から湾奥 (x=x2)まで進むに要する時間aを単位としてはかり,風の継続時間は,T=2\pi\/\omega\=naとする.湾口では常に水位変動がないと仮定すると,\tau\に対して静的な釣合状態を保つ湾奥水位 \eta\ は,海底摩擦がなく湾奥で完全反射の条件が満されているとき,
式(31) [Fig.107_157_02]
であらわされる.
§4にのべたように,防波堤位置における非線型条件の存在のために,パラメータKが重要な量であり,以下の計算では特にことわらないかぎり,\eta\/K=14 の場合が主として計算されている.即ち具体的にいうと,例えばA0/A=1/15,k=0.8とおくと \eta\/h1=0.08であり,(x1+x2)〜40km,h1〜10m 風力係数2×10^{-6} とすると,W (風速)〜30m/sec,\eta\〜80cm の程度である.この模型湾 (長さ40km,深さ10m)ではセイシの基本周期約4時間30分ということになる.
さて,第10図には湾奥に於ける最高波高 \eta\m について \eta\m/\eta\ が,風の吹き続ける時間Tによってどう変化するかが示されている.まず,防波堤のない場合にはT=4aの近傍で,即ち湾の基本振動周期の近くで動力学的な増巾効果が最も著くしく,\eta\m/\eta\ は約1.5となり,T=12aのあたりになるとすでに平衡状態の水面上昇量と等しくなる.このことはすでによく知られていることであつてここで特にのべるまでもない.次に,湾の中央に防波堤のある場合(x1=x2),T=4aのあたりで波高最大値が防波堤のない場合にくらて約半分になつている.これに対してT > 8a となると \eta\m/\eta\〜1 となり,風の周期Tが湾の基本振動周期に比して長いと防波堤の効果はみとめられないことを示している.ここで比較のために,防波堤の開口部がない場合を考えると,湾奥での最大波高は,防波堤のない場合の約1/4程であり,共鳴増巾のおこる風の持続時間Tはaのあたりである.
第11図には防波堤が湾中央 (x1=x2) にある場合について,防波堤位置の直前及び直後の最高水位 \eta\max と風の持続時間Tとの関係が示されている.これによると,T=4aでは,外側最高水位がやや大きく,内側最高水位がやや小さくなることがみとめられるが,T>6a 程度になるとその差が小さくなり \eta\max/\eta\ は約0.5である.
第12図には第11図と同様な条件の場合に,防波堤開口部を通る流量の最大値を示してあるが,風の持続時間Tの変化による最大流量Qの変化はゆるやかである.これを,A0/A=1/15,k=0.8,最大風速30m/sec,h=10m,(x1+x2) = 40kmの場合について換算すると,開口部を通過する海水の最大流速は約5ノットとなる.
第13図には防波堤の開口部の広さ(A0/A)の変化によつて湾奥の最高水位がどう変るかが示されている.勿論A0/Aの小さい程防波堤の効果が大きいが,風の持続時間Tが長いときには,防波堤の有効性は小さなA0/Aの範囲にかぎられることが判る.例えば T=12a ではA0/A〜1/15,T=4a ではA0/A〜1/3の程度で,すでに防波堤の効果がほとんどなくなつている.この図では \eta\ の値を一定(80cm)に保ち \eta\/Kの値は変化させてある.
第14図には防波堤の位置が高潮の高さにどういう影響を与えるかを示してある.ここでは動力学的増巾効果の最も卓越すると思われる T=4a をえらんであるが,この場合,防波堤位置が湾の中央よりやや外海寄りにあるとき湾奥波高が最も低く,波高は防波堤のないときの約1/2である.防波堤位置の内側及び外側の最高波高については,予想される通り,防波堤が湾奥に近い程波高が高くなつている.
第15図には,第14図と同様な場合について防波堤開口部を通過する流量が示されている.ごく湾の奥に位置するときを除いて,防波堤の位置による流量の変化は少い.第12図の説明にあつたと同じ例では,開口部を通過する海水の最大流速は約5ノットに達する.
以上を要約すると,
1) 防波堤のない場合には,風の持続時間がほぼ湾の基本振動周期と一致するときに,湾奥の水位上昇が最大 (定常値の約1.5倍) となるが,湾の中央部に防波堤が存在し,その開口部面積A0が全断面積Aの約1/15程度であれば,湾奥の水位上昇は約1/2に減少する.しかしA0/Aが1/3程度まで大きくなると,防波堤の効果は極めて小さくなる.
2) A0/Aが約1/15の場合,風の持続時間が湾の基本振動周期の約3倍程度になると,湾の中央部に位置する防波堤の効果はほとんどなくなる.
3) 防波堤位置が湾奥に近くなると,湾奥水位を低める防波堤の効果は減少する.例えば防波堤と湾奥との距離が湾の全長の約1/10程度になると,T=4a 程度の持続時間をもつ風に対して,防波堤の波高減衰効果がみとめられない.
4) 防波堤の開口部を通過する海水の最大流量は,風の持続時間や防波堤の位置によつてあまり変化しない.湾の全長40km,平均水深10mのところに最大風速30mの風が吹き,A0/A=1/15,k=0.8とすると,開口部での最大流速は約5ノットとなる.
以上のような結論を実際的な問題に適用するときには,かなりの注意が必要である.理論における一次元の仮定や,摩擦の省略以外に高潮の問題では第一に,湾口水位が常に変化しないというのは事実に反するし,風の吹き方は今仮定されたものよりはるかに複雑である.また風の作用は,水深に関係するから,一様水深を仮定している今の理論値を現実の高潮の高さにむすびつけることには困難がある.これらを考慮すると,この節の結論は定性的なものとみるべきであろう.
- 幅:2099px
- 高さ:353px
- ファイルサイズ:21.7KB
- 幅:1886px
- 高さ:173px
- ファイルサイズ:10KB
- 幅:1399px
- 高さ:1013px
- ファイルサイズ:43.5KB
- 幅:1020px
- 高さ:979px
- ファイルサイズ:29.1KB
- 幅:1067px
- 高さ:747px
- ファイルサイズ:28.2KB
- 幅:1287px
- 高さ:1020px
- ファイルサイズ:43.1KB
- 幅:953px
- 高さ:1592px
- ファイルサイズ:49.9KB
- 幅:1007px
- 高さ:659px
- ファイルサイズ:25.8KB
§8.むすび
すでに述べて来たように,湾内の潮汐,津波,高潮等に対する防波堤の効果を調べるにあたつて,簡単な一次元モデルで防波堤開口部における非線型効果を考えた理論が,模型実験の結果をかなり良く説明することがわかつた.このことは,現実の湾の地形を考慮した二次元モデルによる数値実験では,局部的な非線型効果を考慮に入れることによつて,かなりいい精度で現実の状態を再現する可能性が強いことを示している.ここでは摩擦の効果について何等の考察をしていないが,摩擦による長波のエネルギーの減衰が重要な問題であることは勿論であり,それについてはさらに潮汐,潮流の観測値を利用して研究を進めることが望まれる.
数値実験と模型実験とを比較してその優劣を論ずることは本文の目的ではないが,研究の目的によつて,また必要とする精度によつて,どちらの方法がより有利であるかがきまるものと思われる.例えば,湾の特性のごく大体を知るには実験を行なう必要がないであろうし,複雑な地形の湾や,海岸附近の波の行動を知るには現在のところ大規模な模型実験の方が有利であると思われる.
おわりに,この研究において計算及び製図を手伝つていただいた神里秀子嬢に感謝する.
湾水振動におよぼす防波堤の効果 Effects of a Breakwater on the Oscillations of Bay Water. By Kinjiro KAJIURA, Earthquake Research Institute.
The effects of a large breakwater with a constricted opening on the oscillations of bay water as a whole are examined on the basis of a very simple one-dimensional mathematical model. The bay is separated into two parts by a breakwater with a narrow opening and the dynamical condition at the opening of the breakwater is assumed to follow the wellknown relationship in hydraulics between the volume flux through the opening and the difference of water levels across the breakwater.
Several different cases such as tidal oscillations, invasion of tsunamis, and storm surges are considered and the theoretical results are compared with those of model experiments as far as possible. It appears that a simple theoretical model presented here can explain with satisfaction the observed change of wave height at the head of the bay with respect to the change of the period of forcing functions and/or the area of the opening, as long as the opening is small. Thus, the most important factor in this kind of phenomena should be the following non-dimensional parameter:
B ; for tidal oscillations,
[Fig.107_161_01] ; for tsunamis and storm surges.
For details, see§4 and §3respectively.
The amplitude and phase changes of the tidal oscillation inside the bay can be estimated as a function of B from Fig.2. The effect of a breakwater on invading tsunamis of the solitary type is shown in Fig.9. The effectiveness of a breakwater on storm surges is seen in Figs.10-15. From these figures it is evident that the period of the oscillating water body in relation to the incoming wave period or the given period of external forces seems to be the decisive factor in determining the effectiveness of the barrier, in addition to the geometry of the bay, the area of the opening and the wave amplitude. The present theory is applicable to the cases where the opening is sufficiently small so that the nonlinear effects predominate in the motion of water near the opening.
- 幅:180px
- 高さ:93px
- ファイルサイズ:2.1KB
BULLETIN OF THE EARTHQUAKE RESEARCH INSTITUTE Vol.41 (1963) pp.535-571 33. The Leading Waves of a Tsunami. By Kinjiro KAJIURA, Earthquake Research Institute, (Read March 23,1963. -Rcceived June 27,1963.)
Based on the lillear theory, the decay with distance of the first wave of a tsunami in an infinite sea of constant depth is discussed genera!ly. In particular, for the case of a uniform source distributed in a rectangular area (horizontal dimension :major axis 2a, minor axis 2b) the decay is approximately proportional to the following power of the distance r, except for a distance close to the source area (for comparison, the case of one dimensional propagation is also shown).
[Fig.107_162_01]
[Fig.107_162_02], t:time,H:depth,g;acceleration of gravity.
For a wave ln the direction of the minor axes of the source,a should be replaced by b.
The ratio of the leading wave heights in the directions of the major and the minor axes varies from b/a to 1 depending on the values of pa and pb,so that the directivity of the leading wave height due to the elongated source distribution disappears at a very long distance from the source area. The time interval between the leading wave crest and the second one increases in proportion to t^{1/3}.
The present theory can also be extended to the dispersive wave train in the later phases of a tsunami for an arbitrary source distribution.
- 幅:2712px
- 高さ:854px
- ファイルサイズ:85KB
- 幅:766px
- 高さ:107px
- ファイルサイズ:8.7KB
1. Introduction
Many theoretical discussions have been advanced, mainly in Japan, to explain the generation of tsunami in water of uniform depth when a portion of the bottolm is dislocated vertically.1),2) Most of these linear theories are, however, connned to a rather limited scope of wave properties because of mathematical difficulties confronted in the elucidation of formal solutions. Thus, apart from the numerical approach adopted by several authors,3),4),5),6) common procedures utilized are either the so-called long wave approximation or, at the other limit, the deep water wave approximation, and the analytical discussions including the intermediate range of wave lengths are quite limited because of the transcendental character of the dispersion curve. For a long distance from the source, however, the behavior of a leading wave can be handled analytically by means of an integral analogous to the Airy integral,7),8),9),10) where the approximation involved is essentially similar to the one used to derive the wave equation near the wave front including the effect of curvature of the water surface, and the general characteristics of waves can be estimated from the analysis of this equation.11),12)
However, there seems to be some misunderstanding of the behavior of the leading wave at a long distance from the source. For example, the discussions of the first wave of a tsunami given by Kranzer and Keller(1959),13) and Takahasi (1961) 14) are not valid because of their incorrect application of the ordinary method of the stationary phase up to the wave where the Airy Integral should be considered, and also Van Dorn’s arguments15) on the amplitude decay of the maximum wave with distance is not accepted as evidence of good agreement between theory and observation because he did not compare the corresponding observcd wave at difrerent locations for which the theory predicts the decay law, and besides no distinction was made of the theoretical results for the one-dimensional and two-dimensional dispersion except for the factor r^{1/2}. Similar deficiency of understanding is obvious in Wilson’s discussion 16) of the amplitude decay and period increase of the leading wave.
From a general point of view, this kind of problem is a part of the general theory related to the Cauchy-poisson wave theory concerning the generation and propagation of water waves, and the fundamental properties of linear waves are considered to be well known 17). However, it seems to be worthhwhile to look into the ploblem anew with expression making use of a time dependent Green’s function. By this approach, the dispersive characteristlcs of generated waves and their relation to the nature of the source can be clearly understood.
In particular, this paper deals with the dependence of the dimension of a elevation (or depression) of the water surface on the dimension of a source just after the instantaneous bottom deformation, the distinction between the one and two dimensional decay of the leading wave height at a long distance from the source, and the directional difference of wave height generated by a non-axially symmetric wave source. The limitation of the long wave approximation for the leading wave is also discussed. Finally, the method is extended to treat the dispersive wave train in the later phases of a tsunami for an arbitrary source distribution.
1) Japanese Organization for Tsunami Investigatlon, The annotated bibliography of tsunamis (1962),(mimeographed report).
2) B.W.WILSON, L M. WEBB, and J. A. HENDRICKSON, 「The nature of tsunamis, their gelleration and dlspersion in water of finite depth,」NESCO Tech.Report No.SN 57-2 (1962),146,Appendix 1,2.
3) K.SANO, and K. HASEGAWA,「On the wave produccd by the sudden depression of a small portion of the bottom of a sea of uniform depth,」Bull.Japan Centr.Met. Obsev., 2 (1915),30.
4) T.MATUZAWA,「On the tsunami accompanying the earthqtlake: Pt.1., Two dimensional problem of incompressible water, Pt.2., Numerical comutation of integral,」 Zisin [ii], 1 (1948), 18-23, [ii],2 (1949),33-36, (in Japancse).
5) T.ICHIYE,「A theory on the generation of tsunami is by an impulse at the sea bottom,」Oceanogr. Soc. Japan,14(1958),41-44.
6) loc.cit.,2).
7) K.SEZAWA and K. KANAI,「On the transmission of tsunamis in a sea of any depth,」Bull. Earthq. Res. Inst 20 (1942), 254-264, (in Japanese).
8) R.TAKAHASI,「On the seismic sea waves caused by deformation of the sea bottom,」 Bull. Earthq. Res. Inst. 20 (1942),357-400,(in JapaIlese).
9) C.ECKART,`「The approximate solution of one-dimensional wave equatlons,」Rev. Mod. Phys., 20 (1948), 399-417.
10) loc.cit.,2).
11) G.H. KEULEGAN and G.W.PATTERSON,Mathematical theory of irrotatinal translation waves,」J.Res., Natl.Bureau Stds.,24(1940),47-101.
12) H. JEFFRRYS and B. S. JEFFREYS, Methods of mathematical physics,(Cambridge Press, Cambridge, England,1956),714.
13) H.C.KRANZER and J.B. KELLER,「Water waves produced by explosions,」J.Appl. Phys.,30 (1959),398-407.
14) R.TAKAHASI,「On some model experiments on tsunami generation,」Proc. Tsunami Hydrofynamics Conf., Univ.Hawaii (1961),(publication pending).
15) W・G・VAN DORN,`「Some characteristics of surface gravity waves in the sea produced by nuclear explosions,」J.Geophys. Res.,66 (1961),3815-3862.
16) loc. cit.,2).
17) J.J.ST0KER, Water waves (lnterscience Publishers Inc., New York,1957),569.
2. Fundamental equations and a time-dependent Green’s fuction
Assume incompressible water of constant depth H and take origin of the Cartesian co-ordinate (x’,y’,z’) at the undisturbed free surface with the vertical axis z’ upwards. The irrotational motion in homogencous water may be expressed by means of a velocity potential \psi\ where the velocitoy vector V’ is given by V’=grad \psi\. For convenience in the later discussions, physical quantities are written in nondimensional form unless otherwise noted by putting the independent variables as foliows:
[Fig.107_165_01]
(quantities with ■ime are to be the original form)
where t is time and g is the acccleration due to gravity. The nondimensional form of the derived quantities are:
[Fig.107_165_02]
where \rho\ is density of water, p’ is the anomaly of the surface atmospheric pressure from the mean value, and the vertical component of V is Written as w.
Within the limit of linear approximation (deformations z’ at the surface and at the bottom are assumed small compared with the wave length \lambda\’ and the depth of water H together with the condition [Fig.107_165_03] so that the boundary conditions are satisfied at the undisturbed surfaces), the kinematic and dynamic conditions at the free surface are,
式(1,2) [Fig.107_165_04]
and the bottom condition is,
式(3) [Fig.107_165_05]
where wb is the assumed bottom velocity corresponding to the bottom deformation and the partial differentiations are abbreviated by letter subscripts of the respective variables.
To find a velocity potential \psi\ satisfying (1),(2), and (3) together with suitable initial conditions, it is advantageous to derive a timedependent Green’s function G which should be a harmanic function in the variable (x,y,z) with a singularity of appropriate character at a certain point (xo,y0,z0) introduced at the time t=\time\ Hence,
[Fig.107_166_01]
and G should be a solution of the Laplace equation,
式(4) [Fig.107_166_02]
satisfying the free surface condition
式(5) [Fig.107_166_03]
and a bottom condition
式(6) [Fig.107_166_04]
Atx,y→\infty\, we require G, Gt, and their first derivatives with respect to space coordinates to be uniformly bounded at any given time t. At the point (x0,y0,z0) we require (G-1/R) to be bounded wllere R^{2}=(x-x0)^{2}+(y-y0)^{2}+(z-z0)^{2}. As initial conditions at the time t=\tau\, we assume
式(7) [Fig.107_166_05]
These conditions determine G uniquely.
Following the similar line described by Stoker18), the Green’s function for the case of three dimensional motion in water of finite depth can be derived: for O > z, z0 > -1,
式(8) [Fig.107_166_06]
where r^{2}=(x-x0)^{2}+(y-y0)^{2} and r^{2}=mtanhm. It is evident from (8) that G is symmetric with respect to (x0,y0,z0) and (x,y,z) and also t and \tau\, i. e.:
[Fig.107_166_07]
Making use of the Green’s formula together with the above Green’s function, we may write
式(9) [Fig.107_167_01]
where dS0=dx0dy0 and integral on the lateral boundary in water vanishes
because of the condition imposed on G.
The integration of (9) with respect to \tau\ from 0 to t and the substitution of thc conditions (1),(2),(5),(6), and (7) for \psi\ and G yield,
式(10) [Fig.107_167_02]
The first integral shows the contribution from the surface condition and the second one from the bottom condition.
At the surface z=0,(8) is simplified to give
式(11) [Fig.107_167_03]
and since the surface elevation \eta\ is given by (2), the substitution of (10) into (2) yields
式(12) [Fig.107_167_04]
where
式(13) [Fig.107_167_05]
and
式(14) [Fig.107_167_06]
or alternately,
式(15) [Fig.107_167_07]
and
式(16) [Fig.107_167_08]
or
式(17) [Fig.107_168_01]
It is clear that (13) is the contribution of the initial velocity and elevation of water at the surface,(14) or(15) is the contribution of the the surface pressure, and (16) or (17) is the contribution from the bottom deformation.
The expression (12) together with (13) to (17) is quite general and in essence, represents the application of the principle of superposition in a linear system starting from a point source solution. Since G can be computed without regard to external conditions at the source, it may be quite suitable for numerical computation.
For certain special source conditions with respect to \tau\, (13) to (17) can be simplified as follows with the aid of the Dirac’s Delta function \delta\ (\tau\)19):
(a) water is initially at rest with initial elevation Hs;
式(18) [Fig.107_168_02]
(b) water is initially at rest and applied pressure at the surface is impulsive at \tau\=0+, namely p=Is\delta\(\tau\);
式(19) [Fig.107_168_03]
(c) water is initially at rest and the deformation of the bottom is completed instantaneously at \tau\=0+ with the total deformation HB, namely wB=HB\delta\(\tau\);
式(20) [Fig.107_168_04]
(d) water is initially at rest and uniform velocity of bottom deformation is given for a time interval 0 < \tau\ < \tau\*,with the total deformation HB,namely [Fig.107_168_05];
式(21) [Fig.107_168_06]
(e) water is initially at rest and impulsive bottom motion is given at \tau\=0+ with no net deformation of the bottom after \tau\>0, namely
[Fig.107_168_07]
where IB is the maximum deformation of the bottom;
After slight modification of (17), we have
式(22) [Fig.107_169_01]
It may be possible to represent other kinds of source conditions in simplified forms too; for exarnple, an atomospheric pressure disturbance such as a pressure jump line moving with the constant velocity can be simplified, but further discusslons will not be attempted here.
The comparison of (a) to (e) shows that the instantaneous deformation of the bottom (c) is analogous to a given initial elevation (a), and the bottom impulse (e) can be treated as a surface pressure impulse (b).
The difrerence lies only in the evaluation of the Green’s function G at the surface (z0=0) or at the bottom (z0=-1) which amounts to the decrease of high frequency components by a factor 1/cosh m for the bottom source. As will be shown later in the evaluation of the Green’s function, the leading wave form for a large distance from the source is determined mainly by component waves of very low frequencies \gamma\ (\gamma\ < \gamma\* < 1) so that 1/cosh m is almost one and (a) and (c); (d) and (e) become identical and furthermore,(d) is reduced to (c) provided \gamma\* \tau\* << 1.
For simplicity, We write,
式(23-26) [Fig.107_169_02]
for the cases of initial elevation Hs, initial impulse Is, sudden elevation of the bottom HB, uniform velocity of the bottom deformation HB/ \tau\*, respectively. The functions P,Q,R,and S are considered to be water waves generated by point sources of the particular characters at \tau\=O and are given by evaluating the corresponding form of the Green’s function (11) as follows:
式(27) [Fig.107_169_03]
式(28) [Fig.107_170_01]
式(29) [Fig.107_170_02]
式(30) [Fig.107_170_03]
where r=(x-x0)^{2}+(y-y0)^{2},and for S(\tau\=\tau\*), t should be replaced by t(=t-\tau\*).
The central problem of the wave generation theory in the present formulation is to evaluate P,Q, R,and S as well as possible. Now, returning from (23) to (26), it is easy to show that these expressions comform to the usual expressions derived on the basis of the FourierBessel transform of a potential function and source conditions from the beginning (Appendix I). The advantage of the present formulation may be seen in the separation of the source distribution and the wave dispersion characteristics of the medium.
The Green’s function G*in the two-dimensional motion (x, z plane) can be derived essentially by the similar method in which the singularity imposed on a fluid should be of the type such as (G*-lnR*) to be bounded. In (8), J0(mr) is then replaced by (cosmx)/m where x^{2}=(x-x0)^{2}. In the application of the Green’s formula, in (9), the line integral should be taken instead of the surface integral and 1/(4\pi\) is replaced by 1/(2\pi\). The subsequent deduction is the same and the result can be expressed by
式(31-34) [Fig.107_170_04]
where
式(35) [Fig.107_170_05]
式(36) [Fig.107_170_06]
式(37) [Fig.107_170_07]
式(38) [Fig.107_170_08]
and for S*( \tau\= \tau\*), t should be replaced by t(=t-\tau\*). It is remarked that (35) to (38) can also be derived from (27) to (30) by integration with respect to y0 from -\infty\ to +\infty\ with the aid of a formula,
[Fig.107_171_01]
19) I.N.SNEDDON, Fourier Transform (McGraw Hill Book Co., Inc., New York 1951), 542.
- 幅:1653px
- 高さ:140px
- ファイルサイズ:12.7KB
- 幅:1460px
- 高さ:560px
- ファイルサイズ:39.3KB
- 幅:427px
- 高さ:114px
- ファイルサイズ:4.7KB
- 幅:2207px
- 高さ:339px
- ファイルサイズ:21.3KB
- 幅:2099px
- 高さ:146px
- ファイルサイズ:11.1KB
- 幅:1121px
- 高さ:133px
- ファイルサイズ:10.9KB
- 幅:2281px
- 高さ:133px
- ファイルサイズ:13KB
- 幅:2108px
- 高さ:126px
- ファイルサイズ:10.3KB
- 幅:2048px
- 高さ:140px
- ファイルサイズ:9.8KB
- 幅:2041px
- 高さ:154px
- ファイルサイズ:10.9KB
- 幅:2868px
- 高さ:647px
- ファイルサイズ:69.6KB
- 幅:3035px
- 高さ:153px
- ファイルサイズ:25.4KB
- 幅:2986px
- 高さ:446px
- ファイルサイズ:41.4KB
- 幅:2925px
- 高さ:527px
- ファイルサイズ:50.8KB
- 幅:2879px
- 高さ:387px
- ファイルサイズ:40.5KB
- 幅:2186px
- 高さ:280px
- ファイルサイズ:20.8KB
- 幅:2380px
- 高さ:187px
- ファイルサイズ:16.3KB
- 幅:2452px
- 高さ:227px
- ファイルサイズ:18.3KB
- 幅:2533px
- 高さ:226px
- ファイルサイズ:21KB
- 幅:2439px
- 高さ:234px
- ファイルサイズ:15.3KB
- 幅:2539px
- 高さ:247px
- ファイルサイズ:22.2KB
- 幅:2279px
- 高さ:160px
- ファイルサイズ:14.6KB
- 幅:2279px
- 高さ:174px
- ファイルサイズ:15KB
- 幅:2366px
- 高さ:147px
- ファイルサイズ:15.4KB
- 幅:453px
- 高さ:127px
- ファイルサイズ:5.3KB
- 幅:2533px
- 高さ:160px
- ファイルサイズ:19.5KB
- 幅:707px
- 高さ:233px
- ファイルサイズ:8.8KB
- 幅:2347px
- 高さ:160px
- ファイルサイズ:14.6KB
- 幅:2507px
- 高さ:1073px
- ファイルサイズ:77.4KB
- 幅:2153px
- 高さ:267px
- ファイルサイズ:19.9KB
- 幅:1759px
- 高さ:134px
- ファイルサイズ:8KB
- 幅:1739px
- 高さ:133px
- ファイルサイズ:8.1KB
- 幅:2346px
- 高さ:247px
- ファイルサイズ:23.3KB
- 幅:2486px
- 高さ:893px
- ファイルサイズ:64.4KB
- 幅:2039px
- 高さ:219px
- ファイルサイズ:16.7KB
- 幅:1780px
- 高さ:140px
- ファイルサイズ:8.2KB
- 幅:1752px
- 高さ:127px
- ファイルサイズ:7.7KB
- 幅:2326px
- 高さ:233px
- ファイルサイズ:21.8KB
- 幅:2039px
- 高さ:253px
- ファイルサイズ:22KB
3. Surface elevation at the initial time (t→0) for the case of a sudden deformation of the bottom
For t→0,(29) may be witten as
[Fig.107_171_02]
The expansion of 1/coshm in terms of exponential functions and the integration term by term yield,
式(39) [Fig.107_171_03]
In (39), n can be interpreted as the amount of reflection of the source disturbance at the surface and at the bottom, and the summation of refiection for infinitely many times contributes to the total deformation of the surface.
For one dimensional waves,(37) becomes,
式(40) [Fig.107_172_01]
The initial surface elevations (40) and (39) due to a point source in one and two dimensional waves are shown in Fig.1, which indicates that the elevation is extended over the distance comparable to the depth of water, and no clear-cut wave front is formed.
Talking the origin of the polar-coordinate at the center of an axially symmetric bottom deformation HB, the elevation of the water surface given by (25) at t=0+,r=0 becomes
式(41) [Fig.107_172_02]
For one dimensional waves, the combination of (33) and (40) gives
式(42) [Fig.107_172_03]
particular examples are given below:
α) The uniform deformation of the circular area of the bottom;
式(43) [Fig.107_172_04]
(41) becomes
式(44) [Fig.107_172_05]
and for very small values of a (a << 1),
[Fig.107_172_06]
b) The parabolic deformation of the circular area of the bottom;
式(45) [Fig.107_172_07]
(41) becomes
式(46) [Fig.107_173_01]
where \eta\1/HB is identical to (44) and
式(47) [Fig.107_173_02]
For very small values of a (a << 1),
[Fig.107_173_03]
so that
[Fig.107_173_04]
c)The uniform deformation of the bottom (one dimensional propagation);
式(48) [Fig.107_173_05]
(42) becomes
式(49) [Fig.107_173_06]
and for small values of a (a << 1),
[Fig.107_173_07]
The initial elevation of water surface at the center of the bottom deformation,(44),(46)and (49), is shown in Fig.2, from which it is found that the elevation of the surface at the center rcaches the height of the bottom deformation if the radius of the deformation is about 3 to 4 times the depth of water. On the other hand, for small scale deformation the surface elevation is proportional to the volume of the bottom deformation.
- 幅:1266px
- 高さ:247px
- ファイルサイズ:16.6KB
- 幅:2460px
- 高さ:234px
- ファイルサイズ:21.7KB
- 幅:2067px
- 高さ:1786px
- ファイルサイズ:91.5KB
- 幅:2646px
- 高さ:226px
- ファイルサイズ:25.7KB
- 幅:3026px
- 高さ:233px
- ファイルサイズ:31.1KB
- 幅:2953px
- 高さ:274px
- ファイルサイズ:34.1KB
- 幅:2306px
- 高さ:334px
- ファイルサイズ:24.9KB
- 幅:2406px
- 高さ:233px
- ファイルサイズ:22.4KB
- 幅:1487px
- 高さ:253px
- ファイルサイズ:16.7KB
- 幅:2319px
- 高さ:320px
- ファイルサイズ:26.2KB
- 幅:2013px
- 高さ:194px
- ファイルサイズ:14.1KB
- 幅:2973px
- 高さ:280px
- ファイルサイズ:33.3KB
- 幅:773px
- 高さ:160px
- ファイルサイズ:8.5KB
- 幅:620px
- 高さ:154px
- ファイルサイズ:6.7KB
- 幅:2099px
- 高さ:254px
- ファイルサイズ:20.2KB
- 幅:2380px
- 高さ:220px
- ファイルサイズ:20.2KB
- 幅:1446px
- 高さ:214px
- ファイルサイズ:14.7KB
- 幅:2172px
- 高さ:1420px
- ファイルサイズ:86.4KB
4. Approximate evaluation of the Green’s function
a)Shallow water waves (m << 1, \gamma\=m):
In shallow water, it is usual to adopt the long wave approximation which assumes that the pressure in water is hydrostatic. In the present formulation, the assumption corresponding to the long wave is m << 1 and \gamma\ =m.
The evalution of the Green’s function for this case is rather easy
and the results are,
式(50) [Fig.107_174_01]
and
式(51) [Fig.107_174_02]
If we shart from the beginning on the assumption of a long wave by expanding \psi\ in power series of z and retaining the first order terms only, we arrive at a two dimensional long wave equation with respect
to \zeta\,
式(52) [Fig.107_174_03]
where w is the bottom velocity,\zeta\=\eta\+p, and [Fig.107_174_04]. It should be noticed that S given by (50), where t is replaced by t (=t-\tau\), is nothing but a Green’s function of the wave equation (52) and can be used with advantage for the studies of storm surges20) and tsunamis21).
Near the wave front for a long distance from the source, we may approximate
[Fig.107_175_01]
so that (50) and (51) are reduced to
式(53) [Fig.107_175_02]
式(54) [Fig.107_175_03]
and
式(55) [Fig.107_175_04]
These expressions show that, for a fixed value of (t-r),P,Q,R, and S are all proportional to r^{-1/2} irrespective of the source characteristics and at the front, t=r, the degree of singularity increases from S to P or R and P to Q. However, for a long distance from the source, the usual long wave approximation presented here is not valid clear the wave front because the curvature of the water surface plays a role in the dispersion. In other words, it is necessary to assume \gamma\=m-m^{3}/6 in the neighborhood of r/t \simeq\ 1 and t >> 1.
For one dimensional waves, the same long wave assumption gives
式(56) [Fig.107_175_05]
and
式(57) [Fig.107_175_06]
(56) indicates that one half of the initial surface elevation moves without change of form in the positive and the negative directions respectively.
b) Wave form near the wave front (m < 1, \gamma\=m-m^{3}/6):
For a long distance from the source,(30) may be replaced by
[Fig.107_175_07]
and then transformed into
式(58) [Fig.107_175_08]
It can be shown that the use of the asymptotic expression for J0(mr) results in the error of the order r^{-1} for S. Now, since r, t and \gamma\, m, are positive, the contribution of sin(\gamma\t +m\gamma\-\pi\/4) term to the integral for a long distance is of the order r^{-1}. Thus,(58) may be approximated by
式(59) [Fig.107_176_01]
For the leading wave of a tsunami (r/t \simeq\ 1) for a long distance from the source, it can be shown that the main contribution to S comes from small m so that it is possible to assume \gamma\=m-m^{3}/6 and sinh2m \simeq\ 2m.
Therefore, by replacing m^{3}t/6=u^{6},(59) becomes,
式(60) [Fig.107_176_02]
and
式(61) [Fig.107_176_03]
where Re[z] means the real part of z and
[Fig.107_176_04]
If we neglect u^{6} term in the power of the exponential in (61), we have,
[Fig.107_176_05]
and S is reduced to (53).
Since t is large and the variation of (6/t)^{1/3} with respect to t is small, we may replace the derivative with respect to t or r by the derivative with respect to p, and we may write,
式(62) [Fig.107_176_06]
and
式(63) [Fig.107_176_07]
For one dimensional waves, the asymptotic expression for R* can be derived parallel to the case of two dimensional waves. Thus, for large positive values of t and x,(37) is reduced to
式(64) [Fig.107_176_08]
putting q=p/3=(1/3)(6/t)^{1/3} (x-t), and m=(6/t)^{1/3}v, we have
[Fig.107_177_01]
The integral cal be identified as an Airy Integral Ci(q) where
[Fig.107_177_02]
and
[Fig.107_177_03]
Here, K_{1/3}(x) is a modified Bessel function and J_{1/3}(x),J_{-1/3}(x) are Bessel functions22). (64) is essentially similar to the integral discussed by Eckart23), and Hendrickson24), who treated the asymptotic behavior of the leading wave of the one dimensional tsunami generated by an initial elevation and a bottom deformation of the small scale respectively.
For convenience in later discussions, we define T*(p),
式(65) [Fig.107_177_04]
Then, it follows,
式(66) [Fig.107_177_05]
式(67) [Fig.107_177_06]
and
式(69) [Fig.107_177_07]
Comparing S and S*,R and R*(P and P*), Q and Q*, it is evident that, even if we take the factor r^{-1/2} due to geometrical spreading effect out of consideration, the decay law of amplitude with time for the leading wage is different for the one dimensional and two dimensional waves. Furthermore, the decay laws of the leading waves for the initial elevation and initial impulse are different. A mathematical reason for these differences lies in the fact that, in the evaluation of the integral, the factor in front of the oscillatory term, say, the amplitude spectrum for small values of m plays a deciding role.
For example, the decay law of the leading wave amplitude for a long distance from the source of the type of initial elevation or the sudden deformation of the bottom; namely the variation of the first maximum values of R and R* are (rt)^{-1/2} and t^{-1/3} respectively. The decay law of (rt)^{-1/2} for two dimensional waves was first noticed by Takahasi 25) but was somehow abandoned in his later paper26), and the decay law of t^{-1/3} for one dimensional waves is taken for granted for two dimensional waves as well 27),28), except for the factor r^{-1/2}. In the Appendix II, the decay law for one dimensional waves is derived by means of the superposition of two dimensional waves to show the difference of the decay laws clearly.
For the leading wave generated by a surface impulse, the amplitude decays proportionally to r^{-1/2}t^{-5/6} and t^{-2/3} for the two dimensional and one dimensional cases, respectively.
In both one and two dimensional waves, the wave form of the leading part of the wave train is completely determined by a parameter p so that, with the increase of t, the crest of the wave is retarded with respect to the reference point moving with the velocity of ordinary long wave [Fig.107_178_01], by a factor t^{1/3} and the surging part in front of this crest spreads outward. Furthermore, the time interval between the first crest and the second one increases proportionally to t^{1/3}.
T,Tp, end Tpp, for two dimensional waves and T*,Tp*, and T*pp for one dimensional waves are shown as a function of p in Fig.3 and Fig.
4 respectively. Numerical values of T(p) are origillally given by Takahasi 29), but the re-computation is carried out by a different method.
The result shows a, slight modification of the Takahasi’s values. Apart from the decay factor already mentioned, the wave forms for the two dimensional waves and the one dimensional waves are qualitatively similar provided the position of the first maximum is a little later for the one dimensional wave than for the two dimensional wave. As will be shown in the later section, the wave form for the leading wave of a tsunami at a very Iong distance from the source is considered to be well represented by T, and Tp* for the cases of the initial surface elevation or the bottom deformation and by Tpp and Tpp* for the case of the initial surface impulse. From the figures, it is found that for the one dimensional wave started from the initial elevation or the bottom deformation (Tp*), the leading wave has the maximum height for a very long distance from the source, but for waves started from the initial impulse (Tpp*), this is not the case, and in many cases of practical interest, the leading wave may not be recognized because of its low amplitude and long wave length. For the two dimensional wave (Tp, Tpp), the leading wave may not be the wave with the maximum height for very long distances from the source area, and the second or the later crest will have the maximum height. The intervals between the first maximum and the second one are the same for all Tp,Tp*,Tpp,Tpp*, and the numericcal value in terms of p is 5.7.
Qualitatively similar conclusions concerning the change of the wave form can be obtained on the basis of the one dimensional long wave equation including the effect of curvature of the water surface. According to Keulegan and Patterson 30), the velocity \omega\ of propagation of an element of volume of an intumescence satisfies the relation,
[Fig.107_181_01]
and the wave equation including the second order terms is given by
[Fig.107_181_02]
provided (3/2)\eta^{2}\ << (1/3)\eta\xx or [Fig.107_181_03] (\eta\0:wave amplitude,\lambda\’:wave length, H:depth of water). This wave equation is essentially similar to the assumption \gamma\=m-m^{3}/6. The combination of these two equations and the integration with respect to x give,
[Fig.107_181_04]
From this relation for the velocity of propagation of a volume element of an intumesence moving in still water(say \eta\ > 0), we can easily find that the wave front moves faster than the「long wave velocity」 and the first crest (maximum point) is retarded with respect to the point of inflection where \eta\xx=0 and \omega\=1 as shown graphically in Fig.5.
As for the time interval between the first crest and the second, the different approach made by Munk 31) about the period increase of the conservative waves in general may be applicable. A simple solution for the tsunami given by Munk can be further simplified for the case of a constant depth as follows:
[Fig.107_182_01]
and
[Fig.107_182_02]
where \tau\ is the wave period at the time t and the distance x (all quantities are in the non-dimensional form), and a is an arbitrary constant. For a long distance from the source, the time interval between the first and the second crests is approximately equal to t-x and also to the wave period \tau\, so that we may put \tau\ \simeq\ t -x. Thus we have
[Fig.107_182_03]
Furthermore, since x/t \simeq\ 1,we can conclude that the period increase is proportional to t^{1/3}.
The effect of finite duration of the deforming motion at the bottom can be examined by means of (26), which may be written as
[Fig.107_182_04]
where
[Fig.107_182_05]
In terms of p,is replaced by
[Fig.107_182_06]
where
[Fig.107_182_07]
Judging from Fig.4, the variation of T with respect to p in the neighborhood of (Tp)max is almost linear in the interval (p-p0) < 1, so that
[Fig.107_182_08]
Thus the assumption of the instantaneous deformation is valid in the evaluation of the first crest for the deformation with the time interval \tau\* provided (6/t)^{1/3}\tau\* < 1.For one dimensional waves too, the same condition approximately holds.
c) The wave train in the later phase:
The asymptotic solution of (30) or (56) within the limit of applicability of the stationary phase method is given by,
式(70) [Fig.107_183_01]
where
[Fig.107_183_02]
and
式(71) [Fig.107_183_03]
In (71), primes of \gamma\0 indicate differentiation with respect to m and the sufffix 0 shows the values to be evaluated at m=m0, which is a real root of the equation, \gamma\’= \gamma\/t.
Neglecting the time derivative of the slowly varying amplitude compared with that of the carrier wave, we have by differentiating S with respect to t,
式(72) [Fig.107_183_04]
式(73) [Fig.107_183_05]
and
式(74) [Fig.107_183_06]
For one dimensional waves,(38) is reduced to
式(75) [Fig.107_183_07]
where
式(76) [Fig.107_183_08]
and the other functions can be derived straight-forwardly:
式(77) [Fig.107_183_09]
式(78) [Fig.107_183_10]
and
式(79) [Fig.107_183_11]
For the case of deep water waves, i. e. when the wave number m is much larger than unity and \gamma\ is approximated by m^{1/2}, the asymptotic solutions are reduced to well known formulas where
式(80) [Fig.107_184_01]
式(81) [Fig.107_184_02]
式(82) [Fig.107_184_03]
式(83) [Fig.107_184_04]
The frequency and wave number of the individual wave is fixed by \gamma\0=t/(2r), and m0=t^{2}/(4r^{2}) respectively. The decay laws with distance of the amplitude for a fixed wave number are r^{-1} and x^{-1/2} for the two dimensional and one dimensional waves, respectively.
In both one and two dimensional waves, it is evident that for a fixed location the amplitudes of P and Q increase with time but Q increases more rapidly than P because Q=P, and for Q weight is placed on high frequency components which arrive at the fixed location later.
In contrast, R has a maximum at some intermediate time, because, for very high frequency components of waves, the factor 1/cosh m0 dominates and the amplitudes of high frequency components are surpressed by the factor exp(-m0).
20) K.KAJIURA,「A theoretical and empirical study of storm induced water level anomalies,」 Tech. Report,Ref.59-23F,Dep’t of Ocn. and Met., Texas A and M,(1959), 97.
21) l.N. SRETENSKY and A.S. STAVROVSKY,「Computation of the height of tsunami waves along the coast,」Trans. Marine Hydro. Inst.,Acad.Sci.USSR,24(1961),23-43. (Transl. scripta technica, inc., for the A.G.U.)
22) S.MORIGUCHI, et al,Mathemalical Formulas, III (lwanami Book Co., ltd.. Tokyo,1959),231-232,(in Japanese).
23) loc. cit.,9).
24) loc. cit.,2) Appendix 2.
25) loc. cit.,8).
26) loc. cit.,14).
27) loc. cit.,15).
28) loc. cit.,2).
29) loc. cit.,8) Fig.13.
30) loc. cit.,11).
31) W.H.MUNK,「Increase in period of waves travelling over large distance, with application to tsunamis, swell and seismic surface waves,」Trans. Amer. Geopys. Un.,28 (1947),198-217.
- 幅:2320px
- 高さ:520px
- ファイルサイズ:39.3KB
- 幅:2092px
- 高さ:173px
- ファイルサイズ:12.8KB
- 幅:1933px
- 高さ:174px
- ファイルサイズ:11.6KB
- 幅:726px
- 高さ:134px
- ファイルサイズ:8.6KB
- 幅:686px
- 高さ:147px
- ファイルサイズ:6.5KB
- 幅:2220px
- 高さ:146px
- ファイルサイズ:14.9KB
- 幅:2339px
- 高さ:140px
- ファイルサイズ:14.9KB
- 幅:2220px
- 高さ:173px
- ファイルサイズ:15.6KB
- 幅:2153px
- 高さ:413px
- ファイルサイズ:30.4KB
- 幅:1933px
- 高さ:173px
- ファイルサイズ:10.8KB
- 幅:1767px
- 高さ:260px
- ファイルサイズ:23.1KB
- 幅:2300px
- 高さ:413px
- ファイルサイズ:35.1KB
- 幅:2566px
- 高さ:259px
- ファイルサイズ:27.1KB
- 幅:2099px
- 高さ:167px
- ファイルサイズ:15.6KB
- 幅:2293px
- 高さ:239px
- ファイルサイズ:22.2KB
- 幅:627px
- 高さ:167px
- ファイルサイズ:7.6KB
- 幅:1320px
- 高さ:293px
- ファイルサイズ:18.3KB
- 幅:2186px
- 高さ:173px
- ファイルサイズ:16.7KB
- 幅:2046px
- 高さ:153px
- ファイルサイズ:14.3KB
- 幅:2426px
- 高さ:226px
- ファイルサイズ:25.3KB
- 幅:1420px
- 高さ:227px
- ファイルサイズ:16KB
- 幅:993px
- 高さ:180px
- ファイルサイズ:12.3KB
- 幅:2026px
- 高さ:167px
- ファイルサイズ:19.1KB
- 幅:1979px
- 高さ:247px
- ファイルサイズ:17.5KB
- 幅:1986px
- 高さ:140px
- ファイルサイズ:13.1KB
- 幅:2006px
- 高さ:146px
- ファイルサイズ:10.8KB
- 幅:1979px
- 高さ:167px
- ファイルサイズ:13.8KB
- 幅:220px
- 高さ:136px
- ファイルサイズ:3.2KB
- 幅:2653px
- 高さ:3949px
- ファイルサイズ:266.7KB
- 幅:2892px
- 高さ:3992px
- ファイルサイズ:285.1KB
- 幅:560px
- 高さ:147px
- ファイルサイズ:6.1KB
- 幅:813px
- 高さ:173px
- ファイルサイズ:8.5KB
- 幅:460px
- 高さ:126px
- ファイルサイズ:5.2KB
- 幅:639px
- 高さ:146px
- ファイルサイズ:6.7KB
- 幅:686px
- 高さ:294px
- ファイルサイズ:10.6KB
- 幅:393px
- 高さ:154px
- ファイルサイズ:4.2KB
- 幅:866px
- 高さ:140px
- ファイルサイズ:8.9KB
- 幅:573px
- 高さ:273px
- ファイルサイズ:9.2KB
- 幅:1080px
- 高さ:166px
- ファイルサイズ:11.6KB
- 幅:1860px
- 高さ:140px
- ファイルサイズ:16.8KB
- 幅:1727px
- 高さ:134px
- ファイルサイズ:15.3KB
- 幅:1266px
- 高さ:166px
- ファイルサイズ:14.3KB
- 幅:1167px
- 高さ:927px
- ファイルサイズ:36.8KB
- 幅:2446px
- 高さ:160px
- ファイルサイズ:20.1KB
- 幅:734px
- 高さ:173px
- ファイルサイズ:8.7KB
- 幅:2439px
- 高さ:166px
- ファイルサイズ:16KB
- 幅:2166px
- 高さ:140px
- ファイルサイズ:12.1KB
- 幅:2179px
- 高さ:139px
- ファイルサイズ:12.7KB
- 幅:2186px
- 高さ:160px
- ファイルサイズ:15KB
- 幅:2533px
- 高さ:160px
- ファイルサイズ:21.6KB
- 幅:2079px
- 高さ:140px
- ファイルサイズ:13.6KB
- 幅:2186px
- 高さ:159px
- ファイルサイズ:14.1KB
- 幅:2186px
- 高さ:180px
- ファイルサイズ:16.8KB
- 幅:2180px
- 高さ:167px
- ファイルサイズ:17.6KB
- 幅:1839px
- 高さ:120px
- ファイルサイズ:11.4KB
- 幅:1892px
- 高さ:147px
- ファイルサイズ:11.3KB
- 幅:1866px
- 高さ:140px
- ファイルサイズ:12.6KB
- 幅:1826px
- 高さ:146px
- ファイルサイズ:12.6KB
5.Leading wave of a tsunami
Consider the leading wave of a tsunami for a long distance from the source area enclosed by -a < x0 < a, and -b < y0 < b where waves may be generated either by the deformation of the bottom or by the initial elevation of the surface. Taking the origin of the co-ordinate (x, y) at the center of the source area, we may assume [Fig.107_184_05] where
[Fig.107_184_06]
Geometry of the source point
(x0, y0) and the observing point(x, y) is illustrated in Fig.6, where, for simplicity, the observing point is placed on the x-axis, and \xi\ is taken as the distance of the observing point relative to the moving reference point where x=t. Assuming r \simeq\ x, we may write
式(84) [Fig.107_184_07]
where
[Fig.107_185_01]
Thus, p is independent of y0 for an observing point on the x-axis.
From (25) together with (62), waves generated by the bottom deformation are given by
式(85) [Fig.107_185_02]
Since Tp is independent of y0 for a long distance in the x-direction from the source, we may rewrite (85) with the aid of (84) into the form:
式(86) [Fig.107_185_03]
where
[Fig.107_185_04]
In general, we may put HB=O at the outer edge of the source area, p0=pa and -pa, so that the partial integration of (86) gives,
式(87) [Fig.107_185_05]
where
式(88) [Fig.107_185_06]
If we introduce an explicit distribution of HB, U(p*) can be computed as a function of pa and p* and the surface elevation \eta\ is determined from (87).
For small values of pa, Tp does not change significantly near the maximum of -Tp so that we may approximate (86) by
式(89) [Fig.107_185_07]
where
[Fig.107_185_08]
(89) shows that the shape of the source does not affect the height of the crest or trough of the leading wave for a very long distance from the source (pa << 1). The wave height is proportional to the total volume of the original deformation of the bottom and decays approximately proportional to (rt)^{-1/2} which is already expected from the analysis of R. If the total volume V of the deformation of the bottom is zero, the leading wave at a very long distance from the origin decays inversely proportional to a higher power of the time than 1/2.
For one dimensional waves, we may derive similar equations. From (33) together with (66), the wave generated by bottom deformation is given by
式(90) [Fig.107_186_01]
and (90) may be transformed into
式(91) [Fig.107_186_02]
since
[Fig.107_186_03]
For very small values of pa,(90) is reduced to
式(92) [Fig.107_186_04]
where
[Fig.107_186_05]
particular examples are given in the followings:
a) The uniform deformaction of a rectangulacrarea of the bottom;
式(93) [Fig.107_186_06]
For waves in the x-direction, we may put
[Fig.107_186_07]
so that
[Fig.107_186_08]
(89) gives,
式(94) [Fig.107_186_09]
For small values of pa,
式(95) [Fig.107_186_10]
and for very large values of pa,
式(96) [Fig.107_187_01]
since T(p*+pa)→0 for large values of p*+pa. Here,(-Tp)max is located at p*=-0.75 and Tmax is located at p*=pa-2.3.
b) The elliptic deformation of an elliptic area of the bottom;
式(97) [Fig.107_187_02]
within the region enclosed by
[Fig.107_187_03]
For waves in the x-direction, we may write,
[Fig.107_187_04]
so that
[Fig.107_187_05]
(89) gives
式(98) [Fig.107_187_06]
where \alpha\ = p0/pa and T=T(p*-pa\alpha\).
For small values of pa, we have
式(99) [Fig.107_187_07]
and for large values of pa, we cannot reduce (98) into a simple relation as in the case of a rectangular source.
c) The uniform deformation of the bottom (one dimensional propagation);
式(100) [Fig.107_187_08]
We may write,
[Fig.107_187_09]
and
式(101) [Fig.107_187_10]
For the one demensional leading wave according to (101), \eta_{max}\/HB and p* for which \eta_{max}\ is attained are shown as a function of pa in Fig.7, which indicates that \eta_{max}\/HB is considered to be a linear function of pa for small values of pa,say, for pa < 1, so that the leading wave amplitude varies proportional to a and t^{-1/3}. On the other hand, for large values of pa, say for pa > 3, the leading wave height reaches 0.635HB and does not change with the scale a and the time t. The time interval between the reference point (moving with the long wave velocity √gH from the origin) and the arrival time of the leading wave crest given by p* implies that the crest is retarded with respect to the reference point as the wave travels further, showing the velocity of the crest movement to be smaller than √gH.
For the two dimensional Ieading waves according to (94) and (98), Umax/(2bHB) and p* for with U becomes maximum are shown in Fig.
8,which indicates a trend with respect to pa which is qualitatively similar to the case of the one dimensional wave. As for the directional difference of the leading wave height, the ratio of the leading wave heights in the x- and y-directions for the same time and the distance is given by
式(102) [Fig.107_189_01]
Thus, it is easy to compute the ratio from Fig.8 if pa and pb are known.
For large values of pa and pb, say pa and pb > 3, the leading wave amplitude decreases as r^{-1/2}t^{-1/6} and the amplitude ratio approaches approximately to b/a. On the other hand, for small values of pa and pb, say pa and pb < 1, the amplitude decreases as (rt)^{-1/2} and the directional difference of the leading wave amplitude disappears.
The leading wave generated by the surface impulse can be discussed along a similar line of argument and the surface elevation derived from (24) and (63) is given by
式(103) [Fig.107_189_02]
where
[Fig.107_189_03]
and for small values of pa, we have
式(104) [Fig.107_189_04]
where
[Fig.107_189_05]
For the one dimensional case, the elevation may be derived from (32) and (67)as follows:
式(105) [Fig.107_189_06]
and for small values of pa, we have
式(106) [Fig.107_190_01]
where
[Fig.107_190_02]
Thus the decay of the leading wage height for the case of the surface impulse is greater by the factor t^{-1/3} than that for the case of the initial surface elevation or the sudden deformation of the bottom. However, the directional difference of the leading wave heights follows a rule similar to the case of the initial surface elevation.
- 幅:580px
- 高さ:126px
- ファイルサイズ:6KB
- 幅:500px
- 高さ:127px
- ファイルサイズ:4.5KB
- 幅:2113px
- 高さ:260px
- ファイルサイズ:21.6KB
- 幅:1373px
- 高さ:613px
- ファイルサイズ:29.1KB
- 幅:1246px
- 高さ:153px
- ファイルサイズ:12.2KB
- 幅:2226px
- 高さ:253px
- ファイルサイズ:22.3KB
- 幅:2313px
- 高さ:247px
- ファイルサイズ:23.1KB
- 幅:1440px
- 高さ:260px
- ファイルサイズ:16.1KB
- 幅:2026px
- 高さ:193px
- ファイルサイズ:16KB
- 幅:2172px
- 高さ:227px
- ファイルサイズ:20KB
- 幅:2240px
- 高さ:187px
- ファイルサイズ:18KB
- 幅:1287px
- 高さ:233px
- ファイルサイズ:16.9KB
- 幅:2125px
- 高さ:280px
- ファイルサイズ:20.4KB
- 幅:2205px
- 高さ:219px
- ファイルサイズ:18.2KB
- 幅:1479px
- 高さ:160px
- ファイルサイズ:13.6KB
- 幅:2145px
- 高さ:173px
- ファイルサイズ:16.6KB
- 幅:686px
- 高さ:247px
- ファイルサイズ:8.9KB
- 幅:2305px
- 高さ:293px
- ファイルサイズ:28.9KB
- 幅:1333px
- 高さ:160px
- ファイルサイズ:11.4KB
- 幅:1039px
- 高さ:154px
- ファイルサイズ:11.5KB
- 幅:2132px
- 高さ:173px
- ファイルサイズ:16.1KB
- 幅:2105px
- 高さ:194px
- ファイルサイズ:18KB
- 幅:2026px
- 高さ:166px
- ファイルサイズ:14.2KB
- 幅:2100px
- 高さ:180px
- ファイルサイズ:16.1KB
- 幅:786px
- 高さ:160px
- ファイルサイズ:8.6KB
- 幅:1133px
- 高さ:140px
- ファイルサイズ:12.7KB
- 幅:1433px
- 高さ:167px
- ファイルサイズ:14.9KB
- 幅:2206px
- 高さ:220px
- ファイルサイズ:18.3KB
- 幅:2179px
- 高さ:167px
- ファイルサイズ:17.3KB
- 幅:2152px
- 高さ:306px
- ファイルサイズ:24KB
- 幅:1120px
- 高さ:147px
- ファイルサイズ:12.2KB
- 幅:2326px
- 高さ:160px
- ファイルサイズ:17.8KB
- 幅:2199px
- 高さ:3245px
- ファイルサイズ:224.6KB
- 幅:2033px
- 高さ:173px
- ファイルサイズ:14.9KB
- 幅:2312px
- 高さ:253px
- ファイルサイズ:22KB
- 幅:573px
- 高さ:220px
- ファイルサイズ:7.2KB
- 幅:2253px
- 高さ:180px
- ファイルサイズ:19KB
- 幅:779px
- 高さ:253px
- ファイルサイズ:10.9KB
- 幅:2119px
- 高さ:220px
- ファイルサイズ:18.6KB
- 幅:2166px
- 高さ:146px
- ファイルサイズ:17KB
- 幅:640px
- 高さ:247px
- ファイルサイズ:8.4KB
6.Dispersion of a wave train in the later phase
Making use of the approximate representation of the Green’s function, it is straight-forward to derive the expression of a wave train in the later phase, originating from an extended source area (Ix0I < a,Iy0l < b).
For a large distance from the source,([Fig.107_190_03]), we may approximate
[Fig.107_190_04]
Therefore, the substitution of (72) into (23), which represents the waves started from an initial surface elevation HS, yields,
式(107) [Fig.107_190_05]
For the special case of an axially symmetric source, HS \equiv\ HS(r0), it is easy to show that
式(108) [Fig.107_190_06]
where
[Fig.107_190_07]
In the derivation of (108), the following formulas are used:
[Fig.107_190_08]
and
[Fig.107_191_01]
In a similar way, the waves originating from the axially symmetric surface impulse, Is(r0), are given by
式(109) [Fig.107_191_02]
where
[Fig.107_191_03]
(108) and (109) are identical to the solutions given by Kranzer and Keller 32) who gave the detailed discussions of the wave characteristics derived from these solutions. However, it may be remarked that their conclusions related to the leading wave should be understood with some reservations because the leading wave should be treated as an asymptotic solution of the Airy phase.
It is noticed by examining (107) that for a source function different from axial symmetry, the computation of the elevation becomes complicated in the polar co-ordinate. On the other hand, in the rectangular co-ordinate for \eta\ in the x-direction,(107) can be integrated with respect to y0 without regard to the sine or cosine term since r0cos\theta\0 may be replaced by x0.
Hence,(107) may be written as
式(110) [Fig.107_191_04]
where
[Fig.107_191_05]
The integrals in the right hand side of (110) are nothing but a Fourier cosine and sine transform of W, and represent the amplitude modulation for the carrier wave. For\eta\ in the y-direction, the integration with respect to x can be performed first irrespective of the sine or cosine term slnce we may put r0cos\theta\0=y0.
Some specific examples are given below:
a) The uniform initial elevation of a rectangular area of the surface;
式(111) [Fig.107_192_01]
We have
式(112) [Fig.107_192_02]
Replacing x by y and a by b, the elevation in the y-direction can be obtained,
式(113) [Fig.107_192_03]
Thus, for the amplitude of the carrier wave, the ratio in the x-and the y-directions is
式(114) [Fig.107_192_04]
If we follow the same wave length m0, the amplitude ratio becomes (bsinm0a)/(asinm0b). In particular, if a and b are very small so that m0a, m0b << 1, the directional difference of wave amplitudes in the x-and y-directions disappears. For deep water waves (\gamma\ \simeq\ m^{1/2}) and for the same distance x=y, the ratio of the maximums of the modulation amplitudes in the x-and y-directions becomes b/a and the arrival times of the corresponding maximum amplitudes are given by
[Fig.107_192_05]
Therefore, in the direction of the shorter axis, say in the y-direction if b < a, the arrival of the modulation maximum is later and the wave length of the carrier wave of the maximum amplitude is smaller than those in the x-direction. In other words, the modulation has a larger amplitude and wave length in the direction of the shorter axis than those in the longer axis.
b) The uniform deformation of an elliptic area of the bottom;
Hs=constant.
within the area enclosed by
式(115) [Fig.107_192_06]
We have,
式(116) [Fig.107_193_01]
and the substitution of (116) into (110) yields
式(117) [Fig.107_193_02]
where the following formula is used:
[Fig.107_193_03]
In the y-direction, the wave train is given by
式(118) [Fig.107_193_04]
It is shown that the above expression is reduced to (108) if a=b, because the Hankel Transform of the circular source with constant HS is given by
[Fig.107_193_05]
Now, the ratio of the amplitude of the carrier waves is given by
式(119) [Fig.107_193_06]
This expression is qualitativery similar to the case of a rectangular source provided the sine term is now replaced by the Bessel function. Therefore the qualitative conclusions regarding the difference of the modulation in the x-and the y-directions are similar.
Similar treatments may be applicable to the cases of bottom deformation or with the initial impulse at the surface. For the case of bottom deformation, the result is obtained simply by replacing HS by HB and adding the factor exp(-m0). For the case of the impulsive generation of waves, the result is obtained by replacing HS by Is, changing the carrier waves from cosine to sine, and multiplying the factor \gamma\0 (\simeq\ m0^{1/2}).
32) loc. cit., 13).
- 幅:560px
- 高さ:100px
- ファイルサイズ:6.7KB
- 幅:673px
- 高さ:120px
- ファイルサイズ:6.4KB
- 幅:2813px
- 高さ:486px
- ファイルサイズ:59.4KB
- 幅:2219px
- 高さ:180px
- ファイルサイズ:18.8KB
- 幅:1379px
- 高さ:240px
- ファイルサイズ:16.6KB
- 幅:1106px
- 高さ:220px
- ファイルサイズ:14.6KB
- 幅:907px
- 高さ:240px
- ファイルサイズ:11.9KB
- 幅:2293px
- 高さ:200px
- ファイルサイズ:19.9KB
- 幅:1300px
- 高さ:247px
- ファイルサイズ:16.3KB
- 幅:2626px
- 高さ:533px
- ファイルサイズ:47.7KB
- 幅:1033px
- 高さ:247px
- ファイルサイズ:13.9KB
- 幅:2386px
- 高さ:266px
- ファイルサイズ:30.6KB
- 幅:2506px
- 高さ:160px
- ファイルサイズ:22.3KB
- 幅:2472px
- 高さ:187px
- ファイルサイズ:23.6KB
- 幅:2693px
- 高さ:160px
- ファイルサイズ:25.2KB
- 幅:727px
- 高さ:160px
- ファイルサイズ:7.8KB
- 幅:1972px
- 高さ:174px
- ファイルサイズ:12.9KB
- 幅:2033px
- 高さ:193px
- ファイルサイズ:16.2KB
- 幅:2473px
- 高さ:180px
- ファイルサイズ:23KB
- 幅:1746px
- 高さ:240px
- ファイルサイズ:19.7KB
- 幅:2479px
- 高さ:186px
- ファイルサイズ:23.1KB
- 幅:1220px
- 高さ:153px
- ファイルサイズ:15.2KB
- 幅:2646px
- 高さ:153px
- ファイルサイズ:25KB
7.Comparison with experimental data
a) One dimensional propagation
Prins 33) investigated waves generated by an initial local elevation or depression of uniform height in a two dimensional model. Within the range of H’sa’^{2}/H^{3} < 1 (H’s:the initial height, a’:the half length of the elevation, H:the depth of water), the leading wave height \eta_{max}\’ is found to be proportional to H’s and the proportionality factor may be read off from Fig.7 of his paper.
Depth of water H: 2.3’
Elevation H’s : 0.1’,0.2’,0.3’
Table 1 shows the comparison between the results of the theory and the observation and it may be said that good agreement is obtained for the distance x’=25’. For x’=5’, the agreement is not as good as for x’= 25’,because the approximation made in the theory is not accurate for a short distance from the source.
b)Two dimensional propagation
Takahasi and Hatori 34) carried out a model experiment for the generation of waves by the deformation of a bottom portion of the elliptic shape. The conditions of the experiment are:
Depth H: 5cm and 17.3cm
Dimension of the source:2a’=90 cm,2b’=30 cm
The final form of the displaced bottom surface is approximately parabolic with the maximum height HB0 at the center, and the duration of the bottom motion is of the order of 0.1 sec. (the bottom motion is simulated by the deformation of a rubber memberance). The results which are of interest in the present discussion are shown in Table 2.
The theoretical values for the maximum elevation at the center are estimated approximately by taking (a’+b’)/2 as the representative diameter of the deformation of the type 2b in Fig.2. The observed values for the ratio of the leading wave heights in the directions of the major and the minor axes are the average over the distance 1m to 4m but the theoretical values are computed for the distance of 4m on the basis of Fig.8 curve Ilb.
The observed elevation at the center is about half of the theoretical value and the ratio of the observed wave heights in the major and minor axes for the distance of 4rn is larger than the theoretical values. The reason for this disagreement between theory and observation may lie partly in the inadequate representation of the theoretical model for the actual experimental model conditions, particularly with respect to the time dependence of the bottom motion, and partly in the uncertainty of the model date, which show considerable scattering when several simulations are recorded.
33) J.E. PRINS, 「Characteristics of waves generated by a local disturbance,」Trans. Amer. Geophys. Union,39 (1956),865-874.
34) R.TAKAHASI, and T. HATORI, 「A model experiment of the tsunami generation from a bottom deformation area of elliptic shape,」Bull. Earthq. Res. Inst.,40 (1962),873-883, (in Japanese).
- 幅:3105px
- 高さ:846px
- ファイルサイズ:110.4KB
- 幅:3086px
- 高さ:819px
- ファイルサイズ:107.6KB
8.Concluding remark
The theory developed in the present paper may be severely restricted in application from the practical point of view, because of the various assumptions made in the course of study, such as 1) the linear approximation in the equations,2) constant depth and no lateral boundary,3) the leading wave at long distances from the source,4) time dependence of the source to be of the Delta function type. Since the area of the tsunami generation lies mainly on the continental slope along the pacific Ocean, the assumption (2) should be removed to have a little more realistic picture of the tsunami. The coastal boundary and the continental shelf produce reflected waves and waves travelling along the boundary, so that the wave-train of the tsunami observed along the open coast will be quite different from that expected from the theory with no boundary.
Moreover, the irregularities of the ocean bottom and the existence of islands would change the wave-form significantly during the course of propagation by refraction and diffraction. Unless these factors are adequately taken into account, it seems impossible to understand the wave train of a tsunami clearly. For example, the prevalent period of about one hour observed along the coast of Japan at the time of the tsunami of the Chilean Earthquake cannot be considered to be of local origin, say the oscillation of the bay and the shelf nearby, but of some distant origin. However, it is unlikely that the dispersion is responsible for the wave train of this long period. The reflected waves along the coast of South America and North America, the boundary waves propagated along the coasts and/or the interference of the refracted waves along the course of propagation of the direct waves might be the cause.
As for the assumptions (3) and (4), the numerical computation with the aid of an electronic computor would remove the restriction. Here, the application of the principle of superposition would be very useful to examine various cases of the source condition.
Lastly, the assumption (1) may be justified for the ordinary cases when the deformation of the bottom is not so large compared with the depth of water and also the lateral extent of the source. However, in shallow seas, the condition may not be satisfied and the non-linear effect comes into play. At the coast, the run-up of the tsunami should be discussed separately.
Acknowledgement
The author wishes to express his appreciation of Prof. Takahasi’s enthusiasm for tsunami research which has stimulated his interest in tsunamis and the undertaking of the present study. He also thanks Miss H. Kamisato for her help in various phases of the study.
Appendix I
For example,(23) may be integrated first with respect to dS0 as follows:
式(I-1) [Fig.107_196_01]
Now, since r^{2} = r^{2} +r0^{2}+rr0cosθ0, where (r,0) are the co-ordinates of the point in question and (r0,\theta0\) those of the source point, in the polar co-ordinate with the origin at the center of the source area, we may expand
式(I-2) [Fig.107_197_01]
Thus,(I-1) is reduced to
式(I-3) [Fig.107_197_02]
where
式(I-4) [Fig.107_197_03]
If Hs(r0,\theta\0 \equiv\ Hs(r0):namely for an axially symmetric source,it is easily shown that
式(I-5) [Fig.107_197_04]
and
式(I-6) [Fig.107_197_05]
- 幅:2740px
- 高さ:254px
- ファイルサイズ:31.7KB
- 幅:2566px
- 高さ:387px
- ファイルサイズ:37.4KB
- 幅:2579px
- 高さ:413px
- ファイルサイズ:43.9KB
- 幅:2866px
- 高さ:319px
- ファイルサイズ:36.8KB
- 幅:2493px
- 高さ:260px
- ファイルサイズ:25.1KB
- 幅:2126px
- 高さ:160px
- ファイルサイズ:12.8KB
Appendix II
From the definition of R* and R,it is evident that
式(II-1) [Fig.107_197_06]
where x0 is put zero for simplicity and y0^{2}=r^{2}-x^{2}.
Writing r and x in terms of p and p* for given time t;
式(II-2) [Fig.107_197_07]
we have
式(II-3) [Fig.107_197_08]
and
式(II-4) [Fig.107_197_09]
Since R is very small for large positive values of p and the contribution of R to R*is confined to small values of p only, we may put Approximately,
式(II-5) [Fig.107_197_10]
And the substitution of (II-4),(II-5), and (62) into (II-1) yields,
式(II-6) [Fig.107_198_01]
Thus, R* is proportional to t^{-1/3}.
- 幅:2319px
- 高さ:226px
- ファイルサイズ:21.5KB
- 幅:2399px
- 高さ:166px
- ファイルサイズ:18.6KB
- 幅:2492px
- 高さ:140px
- ファイルサイズ:19.8KB
- 幅:1872px
- 高さ:147px
- ファイルサイズ:11.5KB
- 幅:2400px
- 高さ:160px
- ファイルサイズ:17.8KB
- 幅:2547px
- 高さ:240px
- ファイルサイズ:26.7KB
33.津波の第一波について (津波の発生,伝播に関する古典的理論) 地震研究所梶浦欣二郎
線型近似の範囲内において,海面あるいは海底に与えられたある種の外的じよう乱に原因してお
こる水の波の問題は古くから研究されているが,ここでは時間に関係する型のグリーソ函数を利用
して,一様な深さの海におこる波の問題を統一的に考える.特に外的じよう乱が時間的に簡単な形
で与えられる場合(瞬間的な変動)には波の表現が簡単になり,波の性質を波源の特性(初期水位,
瞬間的な衝撃,瞬間的な海底変動等)と波源のひろがりによるものとに分けて考えることができる.
今 rを波源中心からの水平距離,tを経過時間とすると,波源から極めて遠方における第一波の
減衰は次の量に比例する.
[Fig.107_198_02]
これでみて判る通り,第一波については,一次元・二次元という伝播の相異,および波源の特性の相異が減衰にも関係している.これに反して,第一波と第二波との峯の時間間隔はすぺてt^{1/3}に比例して伸びる.
一方,波源のひろがりの影響については,2aを波源の大ぎさとすると pa =(6/t)^{1/3}a (tおよびαは水深Hおよび重力加速度gを用いて無次元化した量)というパラメータが重要となり,paの大小によつて第一波の減衰の有様が異なる.pa < 1では,波は波源から極めて遠方と考えることができて,減衰は上述の通りであるが,pa > 3では波源にかなり近いところとなり,減哀は次のようになる.
[Fig.107_198_03]
今,長軸2a,短軸2bという矩形の波源から出る波についてその第一波の波高の方向性を調ぺると,近距離では長軸方向の波高は短軸方向の波高の b/a であるが,遠距離では差がなくなる.
定常位相法の利用できる後続波については,円形以外の波源からのものも簡単に調ぺることができて,たとえば楕円波源の場合には,波高のモジュレーションの長さおよび最大振巾は長軸方向よりも短軸方向の方が大きくなることが示される.
得られた結果を実験値と比較すると,一次元伝播の場合には一致が極めて良いが,二次元伝播の場合にはそれほど良くない.
- 幅:2547px
- 高さ:240px
- ファイルサイズ:26.7KB
- 幅:2147px
- 高さ:360px
- ファイルサイズ:35.2KB
Physical processes Associated with the Kuroshio By Kinjiro KAJIURA Earthquake Reseach lnstitute,University of Tokyo
The Kuroshio is a strong current flowing along the western boundary of the North Pacific Ocean all the year round as a part of the general circulation of the ocean, but exhibits fiuctuations of a wide range of frequencies and size scales.
Physical processes corresponding to these fluctuations are quite complicated and not yet well understood because of our lack of knowledge about the non-linear coupling of various dynamical and thermodynamical processes.
To start a discussion, we need to distinguish the scale of phenomena in time and in space. If we formally separate water motion into the mean and fluctuations, the Reynolds stresses enter into the equations of the mean motion due to the non-linearity of the dynamical equations, and the energy exchange of the mean motion and the fluctuating motion as a whole is governed by the tensor product of the Reynolds stresses and the mean velocity strains. Since the Reynolds stresses are essentially related to the fluctuating motion, it is not possible to evaluate the energy flow from the mean motion alone, unless some assumptions are made.
Indeed, in the equilibrium range of the locally isotropic turbulence, the energy flow due to Reynolds stresses is from large eddies to small eddies and the eddy viscosity is formally related to the 4/3 power of the scale is question. In the oceanic scale, we do not know much about the role of Reynolds stresses and the energy might flow from the fluctuating motion to the mean motion (WEBSTER, 1961),but if we apply the statistical theory, the horizontal eddy viscosity seems to be at most of the order of 10^{3} c.g.s. in a climatological sense.
It is well known that, except for the motion with the period of a few days or less, the quasi-geostrophic balance is maintained in most of the oceanic motions except for the surface Ekman layer where the frictional influence is important and this is the basis of the dynamical calculation of ocean currents. However, the computed pressure field from temperature and salinity is only relative and we need to assume the depth of reference pressure somewhere around 1500 m or so. There is a possibility that the depth of no-motion do not exist and the flow may persist to the bottom of the ocean in the same direction as the surface current. Therefore, at the present state of art, it is very uncertain even to say whether the total mass transport of the Kuroshio changes seasonally or not. In deep water, the velocity might be small but its contribution to the total mass transport could be large because of the large thickness of the layer. Furthermore, if the flow reaches to the bottom, the influence of the bottom topography on the flow pattern is expected to be very large, particularly when the flow leaves the continental rise and passes onto the abyssal plain (WARREN,1963).
In the discussion of the circulation, an important concept is the vertical component Z of the absolution vorticity which is the tendency of water particle to rotate and can be separated into two parts;the rotation relative to the earth and the rotation of the earth itself. The Coriolis parameter f gives the vertical component of the planetary vorticity and the ratio of the relative vorticity to the planetary vorticity is given by the Rossby number U/(fL) where U and L are the representative velocity and scale of the motion respectively. If there is no frictional effect, the so-called potential vorticity Z/H (H;thickness of the barotropic layer) and the energy [Fig.107_200_01] of the barotropic fluid are conserved for an individual water column. In the presence of frictional effects, the potential vorticity of a water column may increase or decrease according to the input vorticity, and the change of the vorticity is furnished by the change of the relative and/or the planetary vorticity or by the change of the thickness H of the column. However, the total vorticity integrated over a closed basin is constant for the steady state. This implies that the total input vorticity from the surface of the ocean by the wind stress curl would be cancelled by the vorticity generated by the frictional effect at the lateral boundary (the bottom stress in the interior of the ocean may be negligible).
In the interior of the ocean, it seems reasonable to assume that the Rossby number is very small so that the input vorticity by the wind stress curl is balanced by the change of the planetary vorticity of a water column by the movement in the north-south direction (SVERDRUP,1947). By the continuity requirement of water mass, water return somewhere to the original position. By doing so the movement must also be adequate to generate vorticity which cancels the input vorticity by the wind stress. The relative vorticity and the transport of relative vorticity integrated over the boundary current as a whole is zero and the generated vorticity along the boundary by the friction is just sufficient to balance the change of the planetary vorticity of the boundary current as a whole. To satisfy these conditions, the flow should be in the western boundary (STEWART,1964).
Although the accurate expression of the friction is unknown, if the eddy viscosity of the order of 10^{7} to 10^{8} c.g.s. is assumed, it is possible to give a reasonable picture of the circulation including the western boundary flow. (MUNK,1950).
However, in the western boundary region, the relative vorticity may become comparable to the planetary vorticity, so that the possibility of the inertially controlled boundary current is investigated by neglecting the frictional effect* (CHARNEY,1955, MORGAN,1956). In homogeneous water with constant depth, the inertially controlled strong steady current along the western boundary can be possible only when the boundary current, is accelerated in the direction of flow, and the interior flow is westwards. For the eastward geostrophic flow in the interior, the possible stationary flow pattern may be wave-like and rapidly oscillating in space(FOFONOFF,1962). Another possibility is that the instability of the flow pattern makes the phenomenon always transient and only in a statistical sense ,the frictionally controlled boundary flow would be achieved with the eddy viscosity of the order of 10^{8} c.g.s..
In a two-layer model of the ocean, too, the westward flow is required in the interior of the ocean to establish the western boundary flow (VERONlS,1963).
Here, the density stratification has effects on the flow characteristics. First of all, the total mass transport of the boundary current is limited by the depth of the upper layer in the interior, so that the volume of the upper warm water may control the boundary transport which is otherwise determined by the wind stress curl in the interior of the ocean. Here, we see the interacting mechanism between the dynamical and thermodynamical processes. However, if the barotropic component of the boundary flow is large, the discussion would take a different course. As to the upper warm water mass in the southern half of the interior of the ocean, the thermal circulation in the interior of the ocean and its relation to the formation of the permanent thermocline is discussed by ROBINSON and STOMMEL (1959), WELANDER (1959), and ROEINSON and WELANDER (1963). They solved the heat equation containing three-dimensional advection as well as vertical diffusion in conjunction with the geostrophic equations of motion, and proposed the importance of vertical upward velocity below the layer of the permanent thermocline.
These discussions together with the actual wind stress distribution indicate that in the southern half of the Pacific Ocean an intense northward boundary flow can be formed. It is mainly inertially controlled and includes a frictional boundary layer in the shoreward. On the other hand, in the northern latitude where the geostrophic interior flow is eastward, the strong inertial flow leaves the coastal boundary and the assumption of the steady state may not be justified.
Here, different physical processes should be considered. Indeed, the Kuroshio forms the boundary of the warm southern water and the cold northern water, and the theory developed for the wind stresses should be modified to include the thermohaline processes. particularly, it is desirable to show how much warm water is given off to the north and how much is recirculated in the gyre by taking the meandering and cut-off phenomena of the currents into consideration. Only in a statistical sense could the mean state of the flow be inferred from the frictional theory of the general circulation.
The abyssal circulation of the ocean, which is the internal mode of motion, can also be interpreted by means of the vorticity concept. If we accept the vertical velocity at the interface of the two-layer ocean model, the flow pattern in the lower layer can be estimated by considering the vorticity change due to the stretching of the water column in the lower layer. For the uniform upward velocity at the interface (STOMMEL,1960), the flow pattern in the interior of the ocean is determined, and the concentrated western boundary flow is required to satisfy the continuity equation. However, in the lower layer, the interior flow is eastwards and the boundary flow cannot be of inertial type. This means that the western boundary flow in deep water, if exists as in a model suggested above, might be non-steady and fluctuate with time, so that only in a climatological sense the frictional boundary flow would be realized. A different approach to the internal mode of motion (convective circulation) is developed by TAKANO (1963) who also obtained the intense flow along meridional boundaries. At any rate, it is important to take the internal mode of motion into consideration in the discussion of the general circulation, and the importance of the topographic control of the motion in deep water should be kept in mind.
To examine the time-dependent motion and the stability of the flow pattern, the first step would be to investigate the wave-like motion. A well-known quasigeostrophic wave is the Rossby wave which moves to the west without change of form if no lateral boundary exists. The extension of the Rossby wave to include the wave travelling obliquely to latitude circles is suggested by FOFONOFF (1962) who also clarified the mechanism of the reflection of these waves at the meridional boundary. If we compare barotropic and barociinic waves in a twolayer model of the ocean with respect to the period wave-length relation, it is found that for a scale of several thousand kilometers, the barotropic wave has a period of several weeks and the baroclinic wave has a period of tens of years, so that it is conceivable that the ocean responds mainly in barotropic mode to the external forces of shorter periods but for the seasonal change, the baroclinic response may become about half of the barotropic response (VERONIS and STOMMWL,1956). However, near the equator, the responding baroclinic wave period becomes about a year so that the possibility of the baroclinic resonance exists. It should be mentioned that the situation requires more careful study because the period of the baroclinic wave depends on the latitude and the assumed density structure of the ocean, and moreover the existence of the lateral boundary plays a very important role (Rossby type waves cannot satisfy the reflection condition along the latitudinal boundary) and also the interaction of the barotropic and baroclinic modes may be expected at the boundary.
If the response of the boundary currents to the seasonal variation of the wind stress over the whole ocean is assumed to be mainly barotropic, the amplitude of the variation of the boundary current is roughly proportional to the amplitude of the variation of the wind stress. However, the position of the axis of the boundary current may (lCHIYE,1951, TAKANO,1965) or may not (VERONlS and MORGAN,1955) change with the change of the wind stress. It may be worth-while to investigate whether the current axis of the boundary current changes its position or not when the transport changes suddenly in a two-layer model of the ocean, in connection with the formation of the cold water mass shorewards.
As for the instability of ocean currents, results of several studies have been published (HAURWITZ and PANOFSKY,1950, STOMMEL: The Gulf Stream,1958, IWATA,1961) but our knowledge of the subject is still very far from satisfactory.
Particularly, we need to clarify the instability of flow in relation to the meander and cut-off phenomena when the strong flow leaves the boundary. Here, it may be mentioned that ROSSBY (1951) stated that it is not unreasonable to assume that the factor controlling the shape and behavior of jets must be fairly independent of their driving mechanism and derivable from quite general dynamic principles. He proposed a model of the Gulf Stream on the basis of the lateral mixing of a jet in which the momentum transport is constant and mass transport increases downstream (ROSSBY,1936) and also discussed the vertical and horizontal concentration of momentum in ocean currents by assuming the constant mass transport and the minimum transport of momentum (ROSSBY,1951, ICHIYE, 1960). These arguments are not exact but illuminate some phases of the dynamical behaviors of ocean currents.
As has been discussed so far, the understanding of the oceanic circulation and, in particular, the Kuroshio is very poor and we are just in the beginning of the formulation of a plausible model of the circulation. As in most of the geophyslcal phenomena, theoretical discussions can not go alone without the guidance of observational facts, because the basic form of the hydrodynamic and thermodynamic Iaws cannot be handled without suitable approximations. Thus, it is important to design observational programme suitable to the particular problem in mind. Here, it is necessary to consider the spectral density of the information compared to those of the so-called noises which seems to be very large in most cases (STOMMEL,1963). Therefore, to sort out desired information from the background of noises, continuous time series of observations for a long period, and/or the rapid survey covering wide areas is very desirable, particularly in the upper layers of the ocean and in the region of strong currents where the fluctuations of oceanographic conditions are supposed to be very large.
* CARRIER and ROBINSON (1962) appeared to have pushed the inertial theory to the extreme and their picture of the general circulation is quite unrealistic.
- 幅:3006px
- 高さ:474px
- ファイルサイズ:51.5KB
References
CARRIER, G.F. and A.R. ROBINSON,1960: On the theory of the wind-driven ocean circulation. J. Fluid Mech.,12,49-80.
CHARNEY,I.G,1955: The Gulf Stream as an inertial boundary layer. Proc. Nat. Acad.
Sci. Wash.,41,731-740.
FOFONOFF, N.,1962: Dynamics of ocean currents. The Seas, Interscience Publishers, New York.
HAURWITZ, and H.A. P.ANOFSKY,1950: Stability and meandering of the Gulf Stream.
Trans. Amer. Geophys. Union,31,723-731.
ICHIYE, T.,1951: On the variation of oceanic circulation (1st paper). Oceanogr. Mag., 3,78-82.
ICHIYE, T.,1960: On cirtical regimes and horizontal concentration of momentum in ocean currents with two-layered system. Tellus,12,149-158.
IWATA, N.,1961: Uber die Instabilitat der Meeresstromung im geschichteten Meer.
Contr. Mar. Res. Lab., Hydro. Office of Japan,2,135-170.
MORGAN, G.W.,1956: 0n the wind-driven ocean circulation. Tellus, 8,301-320. MUNK, W.H.,1950: On the wind-driven ocean circulation. J. Met.,7,79-93.
ROBINSON, A.R. and H. STOMMEL,1959: The oceanic thermocline and the associated thermohaline circulation. Tellus,11,295-308.
BOBINSON, A.R. and P. WELANDER,1963: Thermal circulation on a rotating sphere; with application to the oceanic thermocline. J. Mar. Res.,21,25-38.
ROSSBY, C.G.,1936: Dynamic. of steady ocean currents in the light of experimental fluid mechanics. Pap. Phys. Ocn. &Met., 5(1).
ROSSBY, C.G.,1951: On the vertical and horizontal concentration of momentum in air and ocean currents I. Tellus 3,15-27.
STEWART, R.W,1964: The influence of friction on inertial models of oceanic circulation.
Studies on Oceanography,3-9,(Prof. HIDAKA Anniversary Volume).
STOMMEL, H.,1958: The Gulf Stream. Univ. Cal. Press.
STOMMEL, H.and A.B. ARONS,1960: On the abyssal circulation of the word-ocean-I, Stationary planetary flow patterns on a sphere. Deep-Sea Research,6,140-154.
STOMMEL, H.,1963: Varieties of oceanographic experience. Science,139,566-572.
SVERDRUP, H.U.,1947: Wind-driven currents in a baroclinic ocean: with application to the equatorial currents of the Eastern Pacific. Proc. Nat. Acad. Sci.Wash., 33,318-326.
TAKANO, K.,1962: Circulation generale permanente dans un ocean. Rec.Oceanogr.
Works in Japan,6,59-155.
TAKANO, K.,1965: Periodic variation of the barotropic components of a wind-driven circulation in an ocean. Journ. Oceanogr. Soc. Japan,21,1-5.
VERONIS, G. and G.W. MORGAN,1955: A study of the time dependent wind-driven circulation in homogeneous rectangular ocean. Tellus,7,232-242.
VERONIS, G. and A. STOMMEL,1956: The action of variable wind stresses on a stratified ocean. J. Mar. Res.,15,43-75.
VERONlS, G.,1963: On inertially-controlled flow patterns in a \beya\-plane ocean. Tellus, 15,59-66.
WARREN, B.A.,1963: Topographic influences on the path of the Gulf Stream. Tellus, 15,167-183.
WEBSTER, F.,1961: The effect of meanders on the kinetic energy balance of the Gulf Stream. Tellus,13,392-401.
WELANDER, P.,1959: An advective model of the ocean thermocline. Tellus,11,309-318.
BULLETIN OF THE EARTHQUAKE RESEACH INSTITUTE Vol.42 (1964), pp.147-174 7. On the Bottom Friction in an Oscillatory Current. By Kinjiro KAJIURA, Earthquake Research Institute. (Read Dec.17,1963. -Received Dec.28,1963.)
Abstract
The frictional coefficient C of an oscillating turbulent flow is estimated on the assumption that the eddy viscosity is proportional to the amplitude of the bottom friction velocity and the height above the bottom. The dependence of C on parameters such as the amplitude of the vertically averaged horizontal velocity, the period of the oscillation, the depth of water, and the roughness length z0 (for the case of a rough boundary) or the molecular viscosity \nu\ (for the case of a smooth boundary)is shown graphically by choosing suitable non-dimensional parameters. The frictional coefficient for the case of a laminar oscillatory flow is also discussed.
1. Introduction
In many practical applications of the dynamical equations of motion, such as the case of a numerical experiment on long waves in shallow water, the law of bottom stress is assumed as something like [Fig.107_205_01] where u is the depth-mean velocity in a water column and C is the frictional coefficient. There are several estimates of C in tidal currents by means of the dynamical method (Taylor,1918 ; Grace,1936,1937; Bowden and Fairbairn,1952). On the other hand, the bottom friction is intimately related to the mean velocity profile or the turbulent velocity fluctuations near the bottom and there are some observations which show the logarithmic law of velocity profile near the bottom (Lesser, 1951; Charnock,1959). Furthermore, the turbulent velocity fluctuations in a tidal current are measured by means of an electromagnetic flow meter (Bowden and Fairbairn,1956; Bowden,1961) and the Reynolds stresses near the bottom are computed. From these results, it appears that the frictional coefficient (for depth-mean velocity) of a tidal current is about 1.5〜2.5×10^{-3}. However, it is not yet certain whether the same frictional law holds for a long wave of shorter periods, because it is plausible that the structure of the bottom frictional layer in such an oscillatory current may be different from that of the tidal current.
On the other hand, the frictional law of the steady turbulent flow in pipes and open channels is extensively studied by hydraulic engineers but the frictional law of the oscillatory flow seems to be hardly investigated. Recently, the decay of waves in shallow water has become a subject of coastal engineers and some experimental results are discussed in conjunction with the theoretical estimate based on the condition of a laminar flow (Biesel,1949; Eagleson,1962; Grosch,1962), and the transition from the laminar to turbulent boundary layer in an oscillating flow over smooth and rough bottoms is also studied experimentally (Li,1954; Vincent,1957;Collins,1963).
The present paper is a theoretical attempt, though admittedly not so complete, to find the frictional coefficient C of the oscillatory flow in a fully turbulent state by assuming a suitable relation between velocity shear and stress in water and to examine the dependence of the frictional coefficient on the amplitude and period of oscillating currents as well as to the depth of water.
- 幅:267px
- 高さ:127px
- ファイルサイズ:3.1KB
2. Basic consideration
We consider that the motion is predominantly horizontal in onedirection and water is homogeneous with constant depth. Then, the linearized equation of horizontal motion may be written as
式(2-1) [Fig.107_206_01]
where t is time, x and z are the horizontal and vertical co-ordinates with the origin at the bottom of the water and the z-axis positive upwards, u is the horizontal velocity,\zeta\ is the elevation of the water surface from the undisturbed free surface, zh is the depth of water, and g is the acceleration due to gravity. The tangential stress in the x-direction is given by \rho\\tau\ with \rho\ the density of water.
Now putting formally
式(2-2) [Fig.107_206_02]
we may rewrite (2-1) in the form
式(2-3) [Fig.107_206_03]
In terms of the vertically averaged velocity,(2-3) becomes
式(2-4) [Fig.107_207_01]
where \tau\s and \tau\B are the surface and the bottom stresses respectively, (\tau\s = 0 for the present discussion) and u is the mean velocity defined by
式(2-5) [Fig.107_207_02]
The equation of continuity may be given by
式(2-6) [Fig.107_207_03]
and from (2-4) and (2-5) together with (2-2), it follows
式(2-7) [Fig.107_207_04]
or in terms of U,
式(2-8) [Fig.107_207_05]
The equation (2-7) is the common representation of long waves of small amplitude and if the bottom stress is given in terms of u, the solution can be found under given initial and boundary conditions.
The bottom friction velocity u*B defined by
式(2-9) [Fig.107_207_06]
may be approximated for a periodic motion in time (u*=u*Bcos(\sigma\t+\epsilon\)) by
式(2-10) [Fig.107_207_07]
where
式(2-11) [Fig.107_207_08]
and u*B is the amplitude of u*B (See Proudman; Dynamical Oceanography, §151, 1953).
Now, for the relation between the velocity shear and tangential stress, we resort to the usual formulation of a turbulent flow. On the ground of dimensional analysis, the turbulent eddy viscosity Kz in neutral stability may be put formally proportional to the characteristic velocity of turbulence <v^{2}>^{1/2} and the effective size of the turbulent eddy \Lambda\ such that
式(2-12) [Fig.107_208_01]
and
式(2-13) [Fig.107_208_02]
Here, we assume that the effective size of the turbulent eddy is proportional to the height above the bottom and the characteristic turbulent velocity<v^{2}>^{1/2} is proportional to u*B 1). More specifically, we assume
式(2-14) [Fig.107_208_03]
where k is von Karman’s constant (=0.4) and z0 is the roughness length of a rough surface. Similar assumption for the eddy viscosity is already used in the discussion of the atmospheric frictional layer near the ground (e.g., Ellison,1956). For the case of a smooth surface, z0 should be understood as the thickness of a laminar sub-layer. Approximating the frictional velocity in general 2) by
式(2-15) [Fig.107_208_04]
(2-12),(2-13),and (2-15) give
式(2-16) [Fig.107_208_05]
1) The size of turbulent eddies may decrease near the free surface because of the presence of the free surface z = zh, and the characteristic turbulent velocity may not be constant throughout the vertical column of water but a function of the velocity shear and the height above the bottom as assumed in the Prandtl’s mixing-length theory.
However, a refined distribution of Kz does not seem to be worth while to try in the present crude discussion unless more definite knowledge about the turbulent structure of an oscillatory flow is obtained.
2) Essentially, u* is not the friction velocity in an ordinary sense, but is a quantity defined by (2-15) and coincides with the bottom friction velocity uB* at the bottom. For convenience, we call u* the friction velocity in the present paper.
- 幅:2054px
- 高さ:240px
- ファイルサイズ:15.6KB
- 幅:2013px
- 高さ:233px
- ファイルサイズ:14.1KB
- 幅:1947px
- 高さ:260px
- ファイルサイズ:15.9KB
- 幅:2179px
- 高さ:247px
- ファイルサイズ:17.6KB
- 幅:2039px
- 高さ:280px
- ファイルサイズ:15.9KB
- 幅:2053px
- 高さ:267px
- ファイルサイズ:16.6KB
- 幅:2399px
- 高さ:273px
- ファイルサイズ:23.3KB
- 幅:2373px
- 高さ:260px
- ファイルサイズ:27.3KB
- 幅:2026px
- 高さ:220px
- ファイルサイズ:12.1KB
- 幅:2006px
- 高さ:173px
- ファイルサイズ:10.5KB
- 幅:2039px
- 高さ:254px
- ファイルサイズ:14.3KB
- 幅:1846px
- 高さ:167px
- ファイルサイズ:11.2KB
- 幅:1852px
- 高さ:220px
- ファイルサイズ:12.6KB
- 幅:1846px
- 高さ:173px
- ファイルサイズ:11.7KB
- 幅:1786px
- 高さ:167px
- ファイルサイズ:10.2KB
- 幅:1846px
- 高さ:253px
- ファイルサイズ:15.7KB
3. The frictional coefficient for the case of a turbulent flow over a rough bottom
If the motion is assumed to be periodic in time, we may put
式(3-1) [Fig.107_209_01]
where Re means the real part of the complex quantity. In the following discussion, we use the complex amplitude u* with the prime dropped for simplicity unless otherwise stated. The same rule applies for other dependent variables too.
From (2-3) and (2-16) together with (2-15), the equation for the friction velocity u* is derived
式(3-2) [Fig.107_209_02]
where
式(3-3) [Fig.107_209_03]
Under the stress free condition at the surface:
式(3-4) [Fig.107_209_04]
the solution of (3-2) becomes
式(3-5) [Fig.107_209_05]
where
[Fig.107_209_06]
and ch is determined by
式(3-6) [Fig.107_209_07]
Here,
式(3-7) [Fig.107_209_08]
and Jn(Y) and Nn(Y) are Bessel and Neumann functions of the order n and ch=a + ib. For small values of Y,Zn(Y,ch) can be expressed in series form as follows:
式(3-8) [Fig.107_210_01]
and
式(3-9) [Fig.107_210_02]
where r=0.5772.... (Euler’s constant).
The values of a and b computed numerically from (3-6) are shown in Fig.1 as a function of yh.
For small values of yh, say yh < 1,
式(3-10) [Fig.107_210_04]
and
式(3-11) [Fig.107_210_05]
For large values of yh, say yh > 3,
式(3-12) [Fig.107_210_06]
Therefore, it follows:
式(3-13) [Fig.107_211_01]
and
式(3-14) [Fig.107_211_02]
where H1^{(2)}(Y) is the Hankel function of the second kind. (3-13) shows that for small yh the stress \tau\ decreases linearly with respect to depth from the bottom to the surface, and (3-14) shows that the decrease of stress is exponential for large values of y.
Now, in terms of y,(2-16) may be transformed into
式(3-15) [Fig.107_211_03]
and the substitution of (3-5) into (3-15) yields
式(3-16) [Fig.107_211_04]
where the boundary condition at the bottom such that u=0 at y=y0 is taken into consideration. It can be easily shown that for small values of y (y < 1), the velocity profile in terms of z is given by
式(3-17) [Fig.107_211_05]
In terms of y,(2-5) becomes
式(3-18) [Fig.107_211_06]
and the substitution of (3-16) yields
式(3-19) [Fig.107_211_07]
Since in general y0 is very small and (y0/yh)^{2} << 1,(3-19) is reduced to
式(3-20) [Fig.107_211_08]
where
[Fig.107_212_01]
and
[Fig.107_212_02]
For small values of yh, it follows,
式(3-21) [Fig.107_212_03]
and for large values of yh,
式(3-22) [Fig.107_212_04]
The frictional coefficient C is now defined by
式(3-23) [Fig.107_212_05]
where u is the amplitude of u. 3) Therefore, the substitution of (3-20) yields
式(3-24) [Fig.107_212_06]
Taking (3-21) and (3-22) into consideration, it is found that for yh < 1, the frictional coefficient is almost independent of wave period and amplitude and only a function of zh/z0, or in other words, the steady fiow condition is applicable to the oscillatory flow in this range. On the other hand, for yh > 3 the frictional coefficient is a function of y0
3) Strictly speaking, the instantaneous friction coefficient C should be defined by
[Fig.107_212_07]
but for the periodic motion, the following approximation may be introduced;
[Fig.107_212_08]
only so that C is independent of the depth but dependent on wave period and amplitude.
Combining the two relations,(3-20) and (3-24), C (amplitude of C) can be given as a function of u/(\sigma\zh) with zh/z0 as a parameter as shown in Fig.2a. For small values of u/(\sigma\zh), C increases with decreasing values of u/(\sigma\zh) for a fixed value of zh/z0, and for u(\sigma\zh) larger than, say 10^{2},C stays almost constant. Fig.2b shows the corresponding phase angle I\theta\I as a function of u(\sigma\zh) with zh/z0 as a parameter, and it is seen that the phase angle is noticable only for small values of u(\sigma\zh) where the increase of C is observed.
For a fixed values of u(\sigma\z0) , the increase of zh/z0 is followed by the decrease of C and the increase of I\theta\I as shown Fig.3a and Fig.3b. However, for large values of zh/z0, C becomes constant for a fixed value of u/(\sigma\z0), showing the independence of C on zh. In this range, C can be given as a function of u/(\sigma\z0) as shown in