Wednesday, November 30, 2022

2022年11月30日鰻苗捕撈概況

 2022年11月30日(週三)鰻苗捕撈概況

節錄自<日本養殖新聞>
⭕️瓶頸中的台灣,及正要開始的中國,昨晚皆受強風影響,以捕撈漁獲不佳收場!
【台灣】
宜蘭縣昨晚吹起11級強風,使得船隻、定置網及手撈網的漁獲都不好。整體捕撈量本地加上淡水及桃園,看多者有5公斤,多數認為只有2-3公斤,比昨天更少了。
【中國】
昨晚,福建省霞浦及長樂也因為強風,致使船隻無法出去作業,僅在港灣內捕撈。根據多數業者說,整體捕撈量在10-15公斤之間,其中有一半是霞浦的漁獲。

2022年11月29日鰻苗捕撈概況

 2022年11月29日(週二)鰻苗捕撈概況

節錄自<日本養殖新聞>
⭕️台灣依然抓不到,等待新一波潮水進來。中國福建省,捕撈量二位數慢慢的增加中❕
【台灣】
這幾天,台北白天氣溫高,是否因此影響了北部的捕撈?以宜蘭縣海域為中心的鰻苗漁況,因為鰻苗分散四處,根本抓不到,加上鱸鰻苗所佔比例很高,因此增加了許多捕撈作業上的困難(海邊業者說)。既然出去抓不到苗,漁獲也無法應付油錢的支出,因此出去作業船隻也就持續減少。昨晚,根據業者表示,全部約5公斤上下。
【中國】
過去歷年的經驗,正式漁期都從12月中旬開始。本漁季,從前些日子起漁影就開始出現,福建省長樂、霞浦周邊的試驗性捕撈一直持續進行著。目前依然在特定區域不能越界,昨晚的捕撈量增加,主要的長樂有3-4公斤,霞浦3-4公斤,連江1-2kg,南部的莆田、泉州、漳州、廈門及北邊區域目前還抓不到。昨晚的漁獲成績,根據多數業者表示有10-15公斤、20公斤,甚至20-25公斤,正慢慢增加中。

2022年11月28日鰻苗捕撈概況

 2022年11月28日(週一)鰻苗捕撈概況

節錄自<日本養殖新聞>
⭕️昨晚,台灣的漁獲只剩下個位數,僅5-7公斤!
台灣的鰻苗漁況、從上週開始,受到氣溫升高及季節外的颱風影響,漁獲量一口氣向下掉落。再加上幾天前已過漁季的鱸鰻苗混雜其中,每天都是反客為主的漁獲。結果,3位數100公斤以上的漁獲只出現短暫的時間,在不確定的情況下,昨晚終於掉到個位數的5-7公斤。
昨晚宜蘭海域在抓不到的情況之下,出海作業船隻減少到50艘以下。再加上定置網跟手撈網狀況也不佳,整體只有5-7公斤,另一說法,多一倍的漁獲的說法也有。北部淡水、桃園及南部屏東縣還是不變,依然抓不到。這個捕撈狀況一直持續,使得海邊價及流通價也開始上揚。
中國的消息並不多,昨晚,因為還不是可以捕撈時期,所以出去作業的船隻很少。另有一說,福建長樂、霞浦海域比前一天稍微多一些,增加到二位數的10公斤以上的漁獲量。中國各個海邊,少數持有鰻苗的海邊業者也開始與當地的流通業者有所接觸了。

Sunday, November 27, 2022

2022年11月27日鰻苗捕撈概況

 2022年11月27日(週日)鰻苗捕撈概況

節錄自<日本養殖新聞>
⭕️前一天晚上由於新苗混雜其中,因此昨晚原本期待有好消息的,卻又因為鱸鰻苗突然增加而消失,僅有剩下漁群的鰻苗,漁獲僅10-15公斤,抓不到。
昨晚,全區域感覺只有剩下的漁群,因此只有10-15公斤,以抓不到收尾。正如前天所述,昨晚的宜蘭海域鱸鰻苗甚至比前天還多(當地業者說,佔了全部漁獲的90%),新鰻苗的話題完全消失殆盡。
捕苗船漁獲僅有最低水準的500尾/船。結果,昨晚無法確認是否有新鰻苗群。其它地區如北部淡水、桃園及南部屏東縣水域依然抓不到,魚量不多,整體捕撈量根據多數業者說,在10公斤、10-15公斤範圍內。
中國方面,還是在試驗捕撈的海域中。昨晚福建省長樂及霞浦海域有數艘捕苗船出去作業,漁獲量有5-7公斤及7-8公斤兩種說法。業者說「中國依照往年慣例,要12月中旬才會正式進入漁期,因此不到那個時侯,很難說什麼~」,認為現在說什麼都還為時過早。

2022年11月26日鰻苗捕撈概況

 2022年11月26日(週六)鰻苗捕撈概況

節錄自<日本養殖新聞>
鰻苗群到底跑到哪裏去了? 昨晚漁獲跟前一天一樣只有10公斤,不過其中有混雜一些新苗,這使得前景頓時出現曙光!
有些業者描述海邊業者的聲音「現在宜蘭縣的主要捕撈海域,鰻苗數量很少。其中鱸鰻苗雖然不像昨天那麼多,但還是混雜在一起」,出去作業的船隻也更少了。昨晚出去作業船隻少於100艘,每艘船約500尾左右。船東說、儘管數量這麼少了,其中還是有鱸鰻苗參雜在其中。雖然有抱怨捕撈漁獲量不佳,但對今明兩晚的漁況還是很樂觀的。漁獲量有一部分業者說10-20公斤,但多數認爲在10公斤左右。
鰻苗漁況仍然是以宜蘭縣為中心,而南部屏東縣及北部淡水、桃園海域的情況,承如所知,捕撈漁獲均不理想。
中國,依然在試驗捕撈的指定海域中作業,福建長樂海域,昨晚有幾艘船出去,漁獲量沒變加上其它的漁獲約2-3公斤左右。

2022年11月25日鰻苗捕撈概況

 2022年11月25日(週五)鰻苗捕撈概況

節錄自<日本養殖新聞>
⭕️台灣,抓到大量的鱸鰻稚魚,日本種鰻苗更少了,只剩10-15公斤。海邊業者對接下來的漁況,也開始疑惑了。
⭕️中國的試驗性捕撈,昨晚抓的2-3公斤❕
昨晚,台灣的宜蘭縣不知是否是潮水改變?鱸鰻稚魚突然大量增加,讓人相當驚訝,同時也造成選別工作上的麻煩。整體的捕撈量是逐日的減少,昨晚鱸鰻稚魚約佔整體6成,日本苗部分看多的業者是20公斤左右,多半認為在10-15公斤。當地人說「自從漁期開放以來,捕撈狀況一直都很順遂,也有相當的漁獲量。但是這突如其來大量的鱸鰻苗出現,除了驚訝之外,對於接下來的漁況也開始有所擔心了。原本,在這波大潮中抓得到魚是理所當然的事。但是,不知道是不是大自然的玩笑,進入這波暗夜大潮後,反而抓不到魚。」昨晚又發生這起不尋常的變化,讓我們不禁想看看今晚不知又會有甚麼事情發生。南部屏東縣及北部淡水、桃園的捕撈漁獲量很少。
本刊資訊所累計的範圍及外部情報相比照,現階段台灣的總漁獲量應該在1000-1100公斤。
中國,福建省連江、霞浦、長樂、福州等海域,依然持續試驗性捕撈。昨晚長樂的業者說「試驗性捕撈約有3-5艘漁船會出海作業,每艘船約1000尾」。加上其他還有定置網等,總漁獲量應該有2-3公斤吧?!

2022年11月24日鰻苗捕撈概況

 2022年11月24日(週四)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
🔶似乎只能在僅剩的漁群中進行捕撈?!同時等待新鰻苗群到來❕
昨晩宜蘭縣的漁場,出海的捕苗船也一口氣降到只剩100艘。昨晚出海的捕苗船,漁獲成績跟前一天一樣,最後只能作業到半途就早早收網回港。有些持續到早上的漁船也沒有傳來好消息。每艘船約1000尾左右的漁獲。
南部屏東縣的捕苗船依然沒有出港,南北部的定置網及手撈網的漁獲也都不好的樣子。昨晚整體的捕撈量,根據多數業者的訊息從20公斤至50公斤都有,本刊取中間值20-30公斤。
正在進行漁期開放準備的的中國,福建省長樂周邊試驗性捕撈持續進行中。昨晚有幾艘船出海,每艘船的漁獲都在200尾左右,沒什麼變化。

2022年11月23日鰻苗捕撈概況

 2022年11月23日(週三)鰻苗捕撈概況

節錄自<日本養殖新聞>
雖進入暗夜大潮,但已經連續3天都捕撈不佳(50-60公斤、50公斤、30公斤),潮水僅剩3天對於接下來的漁況,將會如何呢?
不知是否還受地震餘波的影響?台灣的鰻苗漁況,昨晚依舊有250艘船出海作業,但是各船隻都抓得不好,紛紛提早收網回港。捕撈量約500-600尾/船,多數業者說整體漁獲大抵約在30-40公斤之間。
台灣自從漁期開放以來,截至今天為止的捕撈總量,依本刊累計加上相關業者的數據平均之後,約1000-1100公斤,很快就超過一噸了。
北部淡水及桃園,還有南部屏東縣鰻苗漁影都不見了,使得捕撈話題也跟著消失。
中國,福建省長樂海域正式進行試驗捕撈。昨晚有幾艘出去,每艘約150尾,稍稍減少。

2022年11月22日鰻苗捕撈概況

 2022年11月22日(週二)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
儘管在第二波潮水的推瀾之下,鰻苗依然四散各處,因此捕撈量跟前一天一樣只有少少的50公斤。
昨晚宜蘭縣的捕撈狀況,不到是否是受到前一天的餘波盪漾,因此鰻苗還是四散各處? 250艘船出海結果,整體捕撈量約50公斤前後(多數業者說從40公斤、45公斤到40-50公斤或是50-60公斤等多種說法),感覺有更減少了一些。另有一說法,宜蘭的東北方抓不好,南方抓的不錯,不過苗是花蓮游上來的,多數是帶黑筋的苗,推估這一波潮水的鰻苗群可能要結束了。
從第一波潮水每天都是三位數100公斤/天的捕撈量,真的是很期待可以持續下去。但是,後繼的第二波潮水在前天晚上,由於東岸海域發生地震,新鰻苗群受到波及,四散各處、使得捕撈量減半。昨晚鰻苗似乎依然四散各處?因此只能在漁獲不佳的情況下結束。接著迎向即將來臨的暗夜大潮,不知這起地震的意外會不會有後續影響?就讓我們看看今晚的捕撈狀況了。
南部屏東縣,當地業者對於還沒有開始捕撈的狀況顯得不安。定置網跟手撈網一直持續著,但是捕苗船卻遲遲無法出港作業。
 中國,與一開始捕撈狀況就很不錯的台灣相較之下,目前漁況並未進入正式的階段。不過,從幾天前開始福建省長樂就一直進行著試驗性捕撈。昨晚也有2-3艘船出海,每艘都有200-300尾的漁獲,有逐日增加的感覺。昨晚福清、連江、長樂及霞浦地區,合計有1-2公斤的捕撈量

Monday, November 21, 2022

2022年11月21日鰻苗捕撈概況

 2022年11月21日(週一)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
意外!!昨晚台灣東海岸又發生地震,結果漁獲量減半只有50-60公斤!!
昨晚,發生了完全沒有預期的意外!東海岸發生地震,有可能因此造成鰻苗群的分散?
船隻方面,同樣是250艘船出海,每艘船平均漁獲減至1000尾/船。另外,北部桃園、淡水等地的漁獲好像很少的樣子。南部屏東縣,捕苗船還是沒有出海的跡象,僅靠定置網及手撈網維繫著。多數業者皆一致認為昨晚的漁獲量在50-60公斤左右。當地業者說「台灣跟日本一樣都是發生地震相當頻繁的地區,而且會對漁況造成直接的影響,此點令人相當擔心。期待今晚的鰻苗也能跟昨天一樣浮上海面。」
中國的鰻苗捕撈測試持續進行中,根據業者說,一艘船約有二位數的漁獲量跟前一天差不多。

2022年11月20日鰻苗捕撈概況

 2022年11月20日(週日)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
進入漁期的第二波潮水、漁獲量突破單日三位數的100公斤!
進入第二波潮水的台灣鰻苗漁況,除了天氣適合漁撈,潮水狀態也好,漁況從宜蘭的南方到花蓮北方開始集結,新的地區如淡水、桃園海域也開始捕撈了。
捕苗船持續有250艘出海作業,每艘船的漁獲平均約2000尾左右。宜蘭縣東北方的漁場還是受到漂流物及水母的影響,加上可能是鰻苗群的移動,均造成捕苗船的困擾與麻煩。昨晚的漁獲本刊以100公斤為紀錄(多數消息從70公斤開始至100公斤,更甚者110-120公斤都有),是本漁季最多一次的捕撈量。
南部屏東縣因爲漁影依舊稀疏,漁群大小還不到漁船出海的程度,所以目前只有定置網跟手撈網吧?!
下個月即將解禁的中國鰻苗漁況,前天晚上長樂海域有兩艘船出去試驗捕撈,並各自捕獲60尾。昨晚又有一艘出去試捕,抓了80尾,比前天的成績還好。

2022年11月18日鰻苗捕撈概況

 2022年11月18日(週五)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
昨晚的捕撈結果約在40-70公斤之間,其中參雜新鰻苗在內❕
昨晚宜蘭縣海上狀況平穩,潮水不大加上鰻苗四散各處,因此捕苗船間就有抓得好及抓得不好的差異出現。
 業者說「新鰻苗有相當數量開始混雜在其中。昨晚跟前一天一樣捕獲量平均在500尾左右,但是抓到1500-2000尾的船也是有的。」昨晚整體的漁獲量從40-50公斤到50公斤、50-60公斤,甚至60-70公斤都有人說,本刊取中間值55公斤。
 進入第二波潮水後,接下來的狀況會越來越好,每天的捕撈情形及漁獲的數量是否會大幅提升,都是我們需要隨時注意的。南部屏東縣,漁影依然未出現,捕苗船尚未出海作業。

首波鰻苗爆發 全臺已捕獲逾700公斤 岸邊價一尾40元
節錄自<農傳媒>
11月1日本季度鰻苗開捕,首波鰻苗漁獲初步結算至17日止,全臺已逾7百公斤,相較於去年同期只有1百多公斤,今年鰻苗漁況初期狀態極佳,鰻苗群目前游向臺灣東岸,鰻苗漁獲超過9成在宜蘭沿岸捕獲,由於漁季一開始鰻苗收獲頗豐,導致鰻苗現在行情比較差,一尾35至40元。

2022年11月19日鰻苗捕撈概況

 2022年11月19日(週六)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
昨晚,台灣捕撈重心的東北部宜蘭縣,因為漂流物及水母的影響,妨礙了捕撈作業的順暢。不過,南方的花蓮縣卻有不錯的成績,讓整體捕撈量可以維持在50公斤左右(多數業者消息,從40公斤到60公斤之間)。南部屏東縣的漁影依然稀疏,漁船尚未出海進行捕撈作業。只有定置網跟手撈網吧!?
12月才解禁的中國鰻苗漁期,根據相關業者表示「福建省長樂,昨晚有兩艘船出去進行試驗性捕撈,結果都各捕撈到60尾」。

2022年11月17日鰻苗捕撈概況

 2022年11月17日(週四)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
天氣回暖,漁場的鰻苗分散不集中加上潮水的原因吧?!整體漁獲約40公斤,回不到之前的捕撈水準!
漁場變寬廣了,加上進入小潮期,鰻苗分散不集中,捕苗船的捕撈效率不佳,使得不少船隻提早收網回港。昨晚有250艘漁船出去作業,其中繞到花蓮去捕苗的漁船也增加不少。不過,整體的捕撈漁獲並沒有增加。
由於捕苗船捕撈作業時間縮短,所以每艘船的漁獲約在500尾上下,全部船隻加總約40公斤左右(多數業者的消息,有35-40公斤、40公斤、40-50公斤、更甚者50-60公斤都有)。
繞到花蓮捕苗的船,捕回的苗是新苗約840-850公克/5000尾的大小,苗體偏小。有人說,昨晚的漁獲量有20-25%是花蓮抓的。這些花蓮縣的苗可能2-3天就會游向宜蘭縣,所以現在對這些苗的捕撈時機充滿期待感。

Thursday, November 17, 2022

2022年11月16日鰻苗捕撈概況

2022年11月16日(週三)鰻苗捕撈概況
節錄自<日本養殖新聞>
【台灣】
昨晚,台東縣外海發生地震,捕撈量只有前一天的1/3約20多公斤!
正值新鰻苗開始混雜,也是本季將進入正式漁期的第二波之前,在台東縣外海發生2次地震。受到影響的花蓮及宜蘭縣的捕苗船,皆因鰻苗分散四處,捕撈情況不甚理想,而早早收網回港。
相關業者說,昨晚每艘船約200-300尾,漁獲量大幅減少。關於數量部分,諸多說法。從15公斤、20公斤、15-30公斤或是20-30公斤都有,本刊取中間值20公斤。

Tuesday, November 15, 2022

鰻魚的喜、怒、哀、樂


2022年11月15日鰻苗捕撈概況

 2022年11月15日(週二)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
鰻苗漁場從花蓮縣北部一直延伸至宜蘭縣全縣,範圍廣闊!!
昨晚以宜蘭縣全縣及花蓮北部為主軸,強風已經和緩,共計250艘捕苗船(其中有40艘是屏東縣籍的船隻)出海作業。目前是從宜蘭縣東北部延伸至全縣並且花蓮縣北部也開始抓的到苗了。每艘船約2000尾的漁獲量,早上聽漁民說約50公斤左右,之後,當所有船都回港後統計,共計70公斤。新苗混雜在其中,感覺大部分依然是之前漁群所留下來的鰻苗,目前是捕撈範圍擴大了,所以後續情況還要再觀察。
其他如定置網的漁獲也變好了,但是手撈部分卻不理想。南部屏東縣的漁船仍然尚未出海作業。昨晚整體的漁獲狀況,從60公斤到60-65公斤、65-70公斤更甚者70公斤或以上都有人說。本刊取70公斤。
台灣鰻苗期才解禁半個月,想起去年同期漁獲數量(到11月10日約50000-70000尾),這波漁期一開始的漁獲相較之下相當的好,希望可以持續維持下去。從解禁以來這半個月台灣的漁獲總量,諸說紛紜,以本刊每日累計的最大值來看約500-600公斤。

2022年11月14日鰻苗捕撈概況

 2022年11月14日(週一)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
昨晚,天候轉為大浪與強風,只能以大型捕苗船為主要作業中心,漁獲大幅減少,只有25-30公斤!!
昨晚宜蘭縣,天候轉為強風造成海上浪高,導致小型船隻及定置網、手撈網都無法作業。捕苗船約150艘(每艘約800尾)出去作業,除了一小部分船隻外,其餘皆因強風浪高而提早回航。其中小型船隻、定置網跟手撈網更是無法作業。從捕苗船業者傳給商社的消息說,今早全部加算應該有35公斤左右吧!不過,早上9點左右(本地時間)又有消息傳來說只有20公斤。有關捕苗船的其他漁獲消息也都是大幅減少的,昨晚漁獲從20-25公斤或是25-30公斤,看多者40公斤。本刊取中間值25-30公斤。
南部屏東縣鰻苗漁影尚不明確,因此捕苗船還是沒有出去。不過,有相當數量的漁船跑到宜蘭作業。

Sunday, November 13, 2022

2022年11月13日鰻苗捕撈概況

 2022年11月13日(週日)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
昨晚出海船隻增加,新舊鰻苗開始混雜一起,全域捕撈量近70公斤,是漁期以來最多的一次!
當我們認為這波漁群已經開始減少的同時,出海作業船隻增加到200-230艘,新舊鰻苗混雜一起,昨晚的漁獲量接近70公斤。應該是本漁期以來最多的一次吧!
相關業者說「昨晚船隻增加到200-230艘,新鰻苗開始混雜在其中,(比之前抓到的鰻苗規格小),加上其它的捕獲量,看少的有60-70公斤,或是65-70公斤,更甚者有70公斤以上。」
定置網及手撈網因為受到風的影響所以捕撈量一直無法提升。南部屏東縣還是一樣,沒有捕苗船出去作業。

2022年11月12日鰻苗捕撈概況

 2022年11月12日(週六)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
以宜蘭縣為軸心的鰻苗漁況,從東北部往北部移動,莫非高峰期已過了嗎?
台灣的鰻苗漁況,一開始是從宜蘭縣這個單一的漁場展開捕撈,這波漁期一開始的大潮,截至目前捕撈狀況及漁獲,似乎已經無法再向上突破了?!上方的鰻苗也開始有一些帶黑筋的苗混在其中。不過鰻苗捕撈區域還是只限於北部地區。
這一波潮水的鰻苗群,一開始捕撈狀況還不錯,之後就是鰻苗群剩餘的數量,再來就是等待下一波新潮水帶來的鰻苗了。昨晚的捕撈情形,依然有200艘船出去作業,平均每艘船800-1000尾,跟前一天沒什麼差別,只是抓到的鰻苗中,已經開始夾雜帶黑筋的鰻苗在內。多數的業者說,昨晚的漁獲量從40公斤到50-60公斤之間。本刊取中間值50公斤。

2022年11月11日鰻苗捕撈概況

 2022年11月11日(週五)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
這波漁期最早開始的大潮,漁獲量已經從高點開始下滑,準備進入潮水週期了吧?!
 昨晚,宜蘭縣東北部軸心的鰻苗捕撈狀況,計200艘船出海,平均每艘船800-1000尾左右,漁獲已經有開始減緩的趨勢。加計約40-50公斤,就算加上定置網及手撈網也就在45-50公斤左右。有人說這波潮水的高峰期已過,準備進入循環週期了。其他的消息則有40-60公斤,沒有很大的差距。截至目前為止台灣的總捕撈量累算下來約計300公斤左右。

2022年11月10日鰻苗捕撈概況

 2022年11月10日(週四)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
天氣轉暖,全域的捕撈量也稍稍減少來到50公斤。
昨晚,依然以宜蘭縣為主軸,東北地區共200艘捕苗船出海作業。因為氣溫上升,相關業者說「預估到中旬氣溫都偏高,因此對於捕撈漁獲應該會開始有所影響」。結果,昨晚每艘船約1000尾/公斤,稍稍減少。全區域包含漁獲也稍有減少的定置網及手撈網,全區域約50公斤左右。多數鰻苗業者認為從40-60公斤之間,本刊取中間值50公斤。
南部屏東縣依然不見鰻苗蹤影,尚未出去作業

Wednesday, November 09, 2022

2022年11月9日 鰻苗捕撈概況

 2022年11月9日(週三)鰻苗捕撈概況

節錄自<日本養殖新聞>
【台灣】
宜蘭縣的鰻苗船,因為近年少見的提早漁獲,因此出海作業數量每日倍增,昨晚也達200艘之多。平均每艘船漁獲有1000-2000尾,抓多的漁船則有4000尾,雖然漁船數增加但是感覺上漁獲量是分散的。此外漁船、定置網及手撈網等都受到漂流物及海藻等附著的影響,使得捕撈作業並不順利,進而影響漁獲數量。昨晚的捕撈量,消息來源從50公斤開始到50-60公斤,進而60-70公斤都有,本刊取中間值60公斤。
南部屏東縣,還是一樣,漁影尚未現蹤,漁民伺機而行。
相關業者說「本漁季有一個很好的開始,但是從捕撈的狀況來看,有不少人認為這波大潮也許是最高峰期。」一直到本月中旬,天氣還是溫暖,加上潮水等因素,這波好的開始,反倒令人有些擔心。

Tuesday, November 08, 2022

盲鰻不是鰻魚

東港約從1998年開始捕盲鰻。盲鰻俗稱「龍筋」,不論是大火快炒韮黃,或是香酥鹽烤、蜜烤三杯、麻油快炒、薑片清湯,吃起來很有嚼勁,脆脆的,相當美味可口。盲鰻黏液清洗時不易清乾淨,有的人使用啤酒洗。
盲鰻(Myxinidae),名字裡有個“鰻”字,但並不是鰻魚。它最特別的技能,大概就是遭遇襲擊的時候,可以在0.4秒內釋放出大量黏液,充滿捕食者的嘴和鰓,讓對方難以呼吸。這樣一來,鯊魚就算再捨不得,也只能把獵物放生。

一隻盲鰻身上,大約有100個黏液腺(slime gland)負責釋放黏液。起初,它們分泌的黏液可能不到5毫升,連小小的一個茶匙都盛不滿。但在受到攻擊的0.4秒過後,幾毫升的黏液就可以變成滿滿一桶史萊姆,比原來的體積擴大了約10000倍。
當生存遇到壓力時,盲鰻會產生大量的黏液,人們對這個現象可能並不陌生。不過對盲鰻來說,“壓力”究竟是指怎樣一種狀況,又為什麼能讓黏液急劇擴張?
回答這個問題之前,我們先要知道盲鰻製造的是怎樣的黏液。這種可以迅速變大一萬倍來打擊敵人的黏液中,除了水和黏液素等常規成分之外,還有一種十分關鍵的要素——纖維

這些纖維是由盲鰻黏液腺裡的端絲細胞(gland thread cell)製造的,又細又長。平時,這些細絲成捆存在,就像一團一團精心纏好的紗線。而一旦盲鰻被捕食者攻擊,或者感覺到周遭的威脅,線團會在毫秒級的時間裡解散開來。當然,無事發生的時候,線團們也會悄悄地鬆動,只是速度要慢許多。 

Unravelling hagfish slime

    Abstract

    Hagfish slime is a unique predator defence material containing a network of long fibrous threads each ∼10 cm in length. Hagfish release the threads in a condensed coiled state known as skeins (∼100 µm), which must unravel within a fraction of a second to thwart a predator attack. Here we consider the hypothesis that viscous hydrodynamics can be responsible for this rapid unravelling, as opposed to chemical reaction kinetics alone. Our main conclusion is that, under reasonable physiological conditions, unravelling due to viscous drag can occur within a few hundred milliseconds, and is accelerated if the skein is pinned at a surface such as the mouth of a predator. We model a single skein unspooling as the fibre peels away due to viscous drag. We capture essential features by considering simplified cases of physiologically relevant flows and one-dimensional scenarios where the fibre is aligned with streamlines in either uniform or uniaxial extensional flow. The peeling resistance is modelled with a power-law dependence on peeling velocity. A dimensionless ratio of viscous drag to peeling resistance appears in the dynamical equations and determines the unraveling time scale. Our modelling approach is general and can be refined with future experimental measurements of peel strength for skein unravelling. It provides key insights into the unravelling process, offers potential answers to lingering questions about slime formation from threads and mucous vesicles, and will aid the growing interest in engineering similar bioinspired material systems.

    1. Introduction

    Marine organisms present numerous interesting examples of fluid–structure interactions that are necessary for their physiological functions such as feeding [1,2], motion [3], mechanosensing [4] and defence [5]. A rather remarkable and unusual example of fluid–structure interaction is the production of hagfish slime, also known as hagfish defence gel. The hagfish is an eel-shaped deep-sea creature that produces the slime when it is provoked [6]. Slime is formed from a small amount of biomaterial ejected from the hagfish’s slime glands into the surrounding water [7]. The biomaterial expands by a factor of 10 000 (by volume) into a mucus-like cohesive mass, which is hypothesized to choke predators and thus provide defence against attacks (figure 1a) [8]. Such defence mechanisms have been observed in several species of hagfish [8,9].

    Figure 1.
    Figure 1. 

    The secreted biomaterial has two main constituents—gland mucus cells and gland thread cells—responsible for the mucus and fibrous component of slime, respectively [6,10]. The plasma membranes of both kinds of cells shear off when secreted from the slime glands [10,11]. In the present study, we focus on the secreted thread cells, referred to as skeins from here on. Skeins possess a remarkable structure wherein a long filament (10–16 cm in length) is efficiently packed in canonical loops into a prolate spheroid (120–150 μm by 50–60 μm) [7,10] (figure 1b). When mixed with the surrounding water, the fibre (1–3 μm thread diameter) unravels from the skein (figure 1c) and forms a fibrous network with other threads and mucous vesicles. This process occurs on time scales of a predator attack (100–400 ms), as apparent from video evidence [8,12].

    While several studies have revealed the mechanical and biochemical aspects [1317] of slime, little is known about the mechanisms involved in its rapid deployment. More recently, efforts have been made to understand the mechano-chemical aspects of the mucus component, and its swelling and rupture [18,19], but such an approach is yet to be extended to the mechano-chemical processes in the unravelling of skeins. Newby [20] postulated that the fibre is coiled under a considerable pressure and the rupture of the cell membrane allows the fibre to uncoil. However, later studies [12,21,22] have shown that convective mixing is essential for the production of fibres and slime. More recently, Bernards et al. [11] experimentally demonstrated that Pacific hagfish skeins can unravel even in the absence of flow, potentially due to chemical release of the adhesives holding the fibre together, but the time scales observed in their work are orders of magnitude larger than physiological time scales during the attack. Therefore, the key question about the fast time scales involved in this process remains to be answered. Deeper insights into the remarkable process of slime formation will aid the development of bioinspired material systems with novel functionality, such as materials with fast autonomous expansion and deployment. Motivated by the aforementioned experimental studies, our objective in this paper is to investigate the role of viscous hydrodynamics in skein unravelling via a simple physical model, and thus supply a qualitative understanding of the unravelling process.

    The key question we answer here is whether the viscous hydrodynamic unravelling alone can account for the fast unravelling time scales that are observed in physiological scenarios. We hypothesize that suction feeding in marine predators creates sufficient hydrodynamic stresses to aid in the unravelling of skeins and set up the slime network. We develop fundamental insight by considering only the simplest flow fields—uniform flow and extensional flow. Our modelling framework, however, generalizes to complex flow fields that occur in physiological conditions.

    In §2, we present a simple qualitative experiment demonstrating the force-induced unravelling of a hagfish skein. This motivates the model paradigm that follows. Section 3 outlines the problem statement, and we derive the general governing equations. In §4, the equations are solved for skein unravelling in simple flows under different physically relevant scenarios. In §5, we discuss the results in more detail, including the influence of constitutive model parameters for the peel strength, and comment on the qualitative comparisons between the experimental studies and theoretical work.

    2. Unravelling experiment

    To motivate the mathematical modelling, we perform a simple experiment demonstrating the force-induced unravelling of thread from a skein (figure 2; see also electronic supplementary material, video). A skein, obtained from Atlantic hagfish, is held in place by weak interactions with the substrate, and a force is applied to the dangling end using a syringe tip that naturally sticks to the filament. Figure 2 shows the unravelling skein at different time frames. Frame 1 shows the unforced and stable configuration, with no unravelling. Unravelling occurs only when a force is applied from frame 2 onwards. There are events when the thread peels away in clumps, but the orderly unravelling recovers quickly. A minimum peeling force seems required to unravel the thread from the skein. A simple estimate of the minimum peeling force based on weak adhesion (van der Waals interaction) between unravelling fibre and skein gives an estimate of 0.1 μN (see electronic supplementary material, section II).

    Figure 2.
    Figure 2. 

    3. Problem formulation

    To determine if viscous hydrodynamic forces can account for fast skein unravelling, we consider a model of an inextensible slender thread unravelling from a spherical skein. The thread unravels and separates from the skein in response to a local force due to a viscous fluid flow surrounding the connected thread and skein. A schematic is shown in figure 3. Here x(s,t) is the Eulerian (laboratory) coordinate of the centreline of the filament as a function of the Lagrangian (material) thread arclength s0sL(t), with L(t) the time-dependent unravelled thread length. The thread is peeling from the skein at the Eulerian point x(L(t),t)=X(t), which may depend on time if the skein is allowed to move.

    Figure 3.
    Figure 3. 

    3.1. Hydrodynamic force balance

    We assume inertial effects, filament self-interactions, and external Brownian and gravitational forces to be negligible. The fluid dynamics in this situation are described by the Stokes equations. For the most general case of a thread in viscous flow, a local balance of filament forces and viscous forces (using local drag theory for a slender filament) is given by Tornberg & Shelley [23]

    8πμδ(xtu(x,t))=((1+2δ)I+(12δ)s^s^)f.
    3.1
    Here the tangent to the thread is s^, the dynamic viscosity is μ and
    δ=1log(ε2e)>0,withε=rL,
    3.2
    are the slenderness parameter and thread aspect ratio, respectively, with r the thread radius.

    The internal net force per unit length, f, of an inextensible filament is expressed by the Euler–Bernoulli bending theory for an elastic beam, and has both tensile and bending components,

    f(s)=(Txs)s+Exssss,|xs|=1.
    3.3
    Here E is the bending modulus of the thread, T(s,t) is the tension in the filament, and each subscript s denotes one derivative, e.g. xs = ∂x/∂s. The inextensibility condition is |xs| = 1, so s and distance along the thread must always coincide.

    In the spirit of rheology, we consider the response to simple flows to isolate key features of the complex behaviour, obtain analytical results and gain an understanding of the unravelling process. We only consider cases with zero curvature, xss = 0, immersed in one-dimensional flow fields, with the thread aligned with the flow streamlines. Equation (3.1) then reduces to a one-dimensional statement that the component of internal net filament force per unit length f(s) along the streamline (taken as the x direction) is equal to the local viscous drag per unit length,

    4πμδ(xtu(x,t))=f.
    3.4
    Then the one-dimensional form of equation (3.3) with xssss = 0 and the inextensibility condition xs = 1 gives f = −Ts, so that
    Ts=4πμδ(xtu(x,t)).
    3.5
    With xs = 1 and x(L,t)=X(t) we have x=XL+s, where X is the skein position, and thus xt=X˙L˙. We integrate (3.5) from s = 0 to L to find
    T|s=LT|s=0=4πμLδ(X˙L˙1L0Lu(XL+s,t)ds).
    3.6
    We then change the integration variable to x=XL+s, and finally obtain
    TLT0=4πμLδ(L˙X˙+u¯(L,X,t)),
    3.7
    where TL=T|s=LT0 = T|s=0 and u¯(L,X,t) is the average velocity on the filament,
    u¯(L,X,t):=1LXLXu(x,t)dx.
    3.8
    Equation (3.7) expresses the balance between the tension forces at the end of the thread and the drag force on the thread. We shall use this equation to derive a peeling formula for different thread–skein configurations in §4. But first, we need to examine how the thread will peel from the skein to unravel.

    3.2. Unravelling from the skein

    The relationship between R and L, respectively, the radius of the spherical skein and the length of the unravelled thread, is described by volume conservation

    ddt(43πηR3+πr2L)=0L˙=4ηR2R˙/r2.
    3.9
    Here r is the thread radius and 0 < η ≤ 1 is the packing fraction of thread into the spherical skein, assumed independent of R. (In this section, we keep the packing fraction as a variable, but in all later numerical simulations we take η = 1, since the skein is fairly tightly packed.) Explicitly, we have
    R3=R0334(LL0)r2/η
    3.10
    with R0 the initial skein radius and L0 the initial unravelled length. A convenient way of relating R and L is
    R=R0(LmaxLLmaxL0)1/3andLmax:=L0+43ηR03/r2,
    3.11
    where Lmax is the total length of thread that can be extracted and L0 is the initial unravelled length.

    Next, we use a modified form of the work-energy theorem [24] to describe the unravelling dynamics,

    E˙total=(TLFP(V))V,V=L˙,
    3.12
    where E˙total is the rate of change in total energy of the system, TL is the net force (given by equation (3.7)) drawing out the thread at a peeling velocity V, and FP(V) is a velocity-dependent peeling force acting at the peeling site. Neglecting the inertia and changes to the elastic energy of the peeling thread gives
    TL=FP(V),V=L˙.
    3.13

    A natural dimensionless quantity that will determine the dynamics of the unravelling process is given by the ratio of the net viscous drag force on the thread and the resisting peel force, each of which depends on a characteristic velocity U,

    :=FD(U)FP(U).
    3.14

    The functional form of the peeling force, FP(V), in general, is dependent on parameters such as the chemistry of peeling surfaces, velocity of peeling, etc. In the absence of a known functional form for hagfish thread peeling, we use a simple constitutive form of peeling force that includes a wide range of behaviour, given by

    FP(V)=αVm,0m1,
    3.15
    for constant α > 0 and m. Such a power-law form of peeling force has been observed in several engineered and biological systems [2529]. Several other parametric forms of velocity-dependent peeling force exist that are functionally more complex [30,31]. However, to obtain simple and insightful solutions, we use the power-law form defined above. The form (3.15) allows for the limiting case m = 0, a constant peeling force, e.g. to simply counteract van der Waals attractions at the peel site.

    For m > 0, we can rearrange equation (3.13) for the velocity, V=L˙=(TL/α)1/m. Using (3.9), we can then obtain a solution for the case where the tension at the peeling point, TL, is constant,

    43(R03R3)=(TLα)1/mr2tη,
    3.16
    where R0 = R(0). From (3.16), we can easily extract the ‘depletion time’ or ‘full-unravelling time’ tdep by setting R = 0,
    tdep=4ηR033r2(TLα)1/m.
    3.17
    In the next section, we compute this time scale when the skeins are subjected to different hydrodynamic flow scenarios, which cause different time histories of tension, TL(t).

    4. Skein in one-dimensional flow

    Having described the unravelling dynamics in §3.2 for the case of constant tension, TL, we now consider a skein in a hydrodynamic flow where generally TL varies in time as the thread–skein geometry changes during unravelling. In physiological scenarios the flow can arise from the hagfish–predator motion, or the suction feeding of the predator, or a combination of both. To simplify the problem we assume an incompressible flow of the form

    u(x,y,t)=(u(x,t),yux(x,t)).
    4.1
    The thread will be assumed to lie along the x-axis. We solve for the depletion time for four relevant cases: pinned thread in uniform flow (§4.1); pinned skein in uniform flow (§4.2); free skein and thread in extensional flow (§4.3); and free skein splitting into two smaller skeins in extensional flow (§4.4).

    4.1. Pinned thread

    The simplest case to consider is the thread pinned at s = 0 in figure 3, with a uniform flow to the right, u(x,t)=U. This situation can arise in a controlled experiment if the thread is pinned down, or in the physiological unravelling process if the end of the thread is caught in the network of other threads, or stuck on the mouth of a predator.

    The tension in the thread at s=L balances the Stokes drag on the skein of radius RTL=T(L(t),t)=6πμR(u(L,t)L˙). Using (3.13) and (3.15), we obtain the governing equation for unravelling as

    (L˙)m=6πμα1R(L)(u(L,t)L˙).
    4.2
    From (3.9), since L˙>0 (the thread never ‘re-spools’), the unspooling speed satisfies L˙u(L,t), i.e. the thread cannot unspool faster than the ambient flow speed. The radius R(L) is given by (3.10).

    We non-dimensionalize (4.2) using a characteristic length scale R0 and flow speed U, which gives

    (L˙)m=R(L)(u(L,t)L˙),
    4.3
    where L˙=L˙/UR=R/R0 and u(L,t)=u(L,t)/U are the non-dimensional unravelling rate, skein radius and flow rate, respectively. The non-dimensional time scale naturally results from these choices as t=t/(R0/U). The dimensionless quantity  on the right-hand side of (4.3) is given by
    =6πμR0UαUm=6πμR0U1mα1.
    4.4
    This is the ratio of characteristic drag to peeling force, as defined in (3.14). If  is large (e.g. zero resistance to peeling), then (4.2) implies L˙u(L,t), that is, in this drag-dominated limit, the drag force so easily unravels the skein that it advects with the local flow velocity. In the opposite limit of small , we get L˙0 and the skein cannot unravel. Hence, we require ℘ ≫ 1 for a fast unravel time.

    To achieve the criterion ℘ ≫ 1, at a flow of speed U = 1 m s−1 and a skein of initial radius R0 = 50 μm, we require the peeling resistance at this velocity to satisfy FP(1 m s−1) ≪ 1.4 × 10−6 N. The estimated van der Waals peeling force is much lower than this threshold, FvdW ∼ 0.1 μN. At such a flow speed a skein containing 16.7 cm of thread (an upper bound physiological value) will unravel affinely (kinematically matching the flow speed) in roughly 167 ms. This lower bound estimate is commensurate with the rapidity with which hagfish slime is created (100–400 ms).

    In figure 4, we show a numerical solution of (4.2) with a uniform flow for some typical physical parameter values, and assuming a moderately large force ratio ℘ = 10. (Equation (4.2) is an implicit relation for L˙ which must be solved numerically at every time step; it is a differential--algebraic equation rather than a simple ODE [30].) For these parameters, the kinematic lower bound on the depletion time is Lmax/U167 ms, and the numerical value is tdep194 ms.

    Figure 4.
    Figure 4. 

    There is a mathematical oddity where the skein might not get depleted in finite time, depending on the exponent m. To see this, consider a skein close to depletion, L=LmaxUτ, where τ > 0 is small. The equation for τ is

    (τ˙)m=(UτLmaxL0)1/3(1+τ˙),τ˙<0.
    4.5
    Since τ is small and we expect the thread to be drawn out slowly as it is almost exhausted, we take 1+τ˙1. Hence, we have the approximate form
    (τ˙)mCmτ1/3,Cm:=(U(LmaxL0))1/3
    4.6
    for some constant C > 0, with solution
    τ(t)[τ011/3m(113m)Ct]3m/(3m1).
    4.7
    The behaviour of this solution as the skein is almost depleted depends on m. For m > 1/3, the exponent 3m/(3m − 1) in (4.7) is greater than 1, so τ(t)0 as t approaches the depletion time, with τ(tdep)=0 so that L(t) has slope zero when the skein is depleted (as can be seen at the very end in figure 4). We can thus rewrite (4.7) as
    τ(t)[(113m)C(tdept)]3m/(3m1),m>13,ttdep.
    4.8

    For m < 1/3, the exponent 3m/(3m − 1) is negative, but the factor 1 − 1/3m inside the brackets is also negative, so that τ(t) asymptotes to zero as t and the skein never gets fully depleted. In that case, we write (4.7) as

    τ(t)[(13m1)Ct]3m/(13m),m<13,t.
    4.9
    Physically, for m < 1/3 the drag force (∼τ1/3) is decreasing faster than the peeling force (τ˙m).

    In practice, it is difficult to see the difference between m1/3 numerically. The thread appears to get depleted even for m < 1/3 because of limited numerical precision as L approaches Lmax. The symptom of a problem is that the depletion time starts depending on the numerical resolution for m < 1/3. Of course, the skeins in the hagfish slime do not need to get fully depleted to create the gel, so a power m < 1/3 is still applicable. When comparing the different flow scenarios we will explore a range of m and define an ‘effective deployment’ time tdep,50%, when 50% of the thread length is unravelled.

    4.2. Pinned skein

    When the skein is pinned and the thread is free at the other end, the tension arises from hydrodynamic drag on the thread. Such a scenario can arise if the skein is arrested in the network of other fibres or in a mucus network.

    Consider a free thread ending at s = 0 and a pinned skein at s=L(t) (figure 3), so that the Eulerian skein position X is fixed and is thus not a function of time. Unlike the pinned thread case in §4.1, where a shrinking skein led to a decreasing drag, here the tension increases with time as the extended thread provides more drag.

    We formulate the problem by imposing boundary conditions at the free end, T0=T(0,t)=0, and pinned end, x(L(t),t)=X. From (3.7) with T0=X˙=0 and TL=α(L˙)m, the equation for the growth of the thread is

    (L˙)m=4πμα1Lδ(L)(L˙+u¯(L,X,t)).
    4.10
    The slenderness parameter δ depends on L through its definition (3.2). Because the thread extends to the left in figure 3, we must have u¯(L,X,t)<0 to avoid unphysical respooling. The pinned thread equation (4.2) and the pinned skein equation (4.10) have a very similar form, though the drag in the former (R(L)) decreases with L and that in the latter (Lδ(L)) increases with L.

    Using a characteristic velocity U and a characteristic length scale L0, we obtain the non-dimensional form of (4.10) as

    (L˙)m=Lδ(L)(u¯(L,X,t)+L˙),
    4.11
    where L˙=L˙/UL=L/L0 and u(L,t)=u(L,t)/U are the non-dimensional unravelling rate, unravelled length and flow rate, respectively. The natural dimensionless force ratio (3.14) is
    =4πμL0U1mα1.
    4.12
    This differs from  in (4.4) by replacing R0 with L0. It is sensible in this pinned skein case to use the initial thread length L0, since drag on the thread controls the unravelling rate.

    Figure 5 shows a numerical solution of (4.10) for our reference parameter values using a constant velocity field, u(L,X,t)=U=1 m s−1. As before, the lower bound on the depletion time is Lmax/U167 ms, and now the numerical value is tdep226 ms. This is slower than what we observed in the pinned thread case (tdep194 ms), but here we are using the much smaller force ratio ℘ = 1/2. This shows that the pinned skein case can unravel almost as fast as the lower bound for a much smaller value of , since the drag on the thread increases with L, as reflected by the accelerating speed L˙ in figure 5. This is in contrast to the deceleration in figure 4 for the pinned thread, where drag decreases as the skein radius diminishes.

    Figure 5.
    Figure 5. 

    4.3. Free skein and thread

    In the previous two cases, we took either the thread or skein to be pinned; here we consider the case where neither is pinned, and both are free to move with the flow. This scenario is possible if both the thread and the skein are not stuck to the existing fibrous network, or at the beginning of the slime formation when none of the skeins have unravelled to a significant extent.

    The force at the peeling point x(L(t),t)=X(t) is then determined by the balance of two forces: Stokes drag on the spherical skein, F1=6πμR(u(X,t)X˙), and drag on the thread, F2=4πμLδ(L˙X˙+u¯(L,X,t)). The latter is obtained from (3.7) with TL=F2 and T0 = 0. Since both the skein and thread are free and we have neglected inertia, F1 + F2 = 0, which we can use to solve for X˙, the velocity of the peeling point in the Eulerian (laboratory) frame. Coupling this with the peel force constitutive model (3.13)–(3.15), the unspooling rate equation is then α(L˙)m=F1=F2. The dynamics of this scenario are governed by the system

    (L˙)m=12πμα1RLδ2Lδ+3R(L˙+u¯(L,X,t)u(X,t))
    4.13a
    andX˙=2Lδ2Lδ+3R(L˙+u¯(L,X,t))+3R2Lδ+3Ru(X,t),
    4.13b
    where u¯(L,X,t) is the thread-averaged velocity (3.8). The velocity (4.13b) for the thread–skein system is the average of a velocity L˙+u¯(L,X,t) arising from drag on the thread and a velocity u(X,t) arising from drag on the skein, weighed by the relative strength of the drags.

    The difference u¯(L,X,t)u(X,t) that appears in (4.13a) implies that adding a constant to the velocity field does not change the unspooling dynamics, as expected since the thread–skein system is freely advected by the flow, and only relative velocities generate drag. Hence, unlike our previous two cases in §§4.1 and 4.2, a spatially varying flow field is required for unravelling. For a linear velocity field u(x,t)=λx, i.e. uniaxial extensional flow with extensional strain rate λ, we have u(X,t)u¯(L,X,t)=λL/2 independent of X, so that we can solve the L˙ equation (4.13a) by itself,

    (L˙)m=6πμα1RLL+(3R/2δ)(12λLL˙).
    4.14
    The mass conservation equation (3.10) then relates R to L, and the slenderness parameter (3.2) relates δ to L.

    To define a characteristic length scale for this problem, should we use R0 or L0 as a length scale? Both are important for the unravelling process to start quickly, but typically L0 is a bit larger than R0. A compromise is to use R0 as the viscous drag length scale and U=λL0 as the velocity scale. The choice of R0 emphasizes the magnitude of the drag on the skein, and λL0 reflects the amplitude of velocity gradients over the longer length L0. We thus obtain the dimensionless form of (4.14) as

    (L˙)m=RLL+(3R/2δ)(12LL˙),
    4.15
    where L˙=L˙/λL0R=R/R0 and L=L/R0 are the non-dimensional unravelling rate, skein radius and unravelled length, respectively. The natural dimensionless number in this case is
    =6πμR0UαUm=6πμR0(λL0)1mα1.
    4.16

    Assuming as before that L˙0 (the thread does not ‘re-spool’), the right-hand side of (4.14) implies L˙λL/2, which gives the constraint that L(t)L0e(1/2)λt. This constraint is the kinematic limit where the thread extends at a rate dictated by the strain rate in the flow. This implies that the depletion time satisfies

    tdep2λ1log(LmaxL0).
    4.17
    In the two pinned cases we considered before, the lower bound on the depletion time was of the form tdepLmax/Uindependent of L0. The lower bound (4.17) depends explicitly on the ratio Lmax/L0, so a very short initial thread length will take a long time to unravel, even if  is large.

    When the thread is almost depleted, the unspooling rate decreases due to the factor of R in (4.14). To see this explicitly, put L=LmaxUτ in (4.14) and assume τ and τ˙ are small,

    (τ˙)m12(UτLmaxL0)1/3LmaxL0,τ1.
    4.18
    This is exactly the same form as (4.6), with a different constant C. We conclude that once again the criterion for finite-time complete unravelling is m > 1/3, as it was for the pinned thread case (§4.1). But as before this is not very physically consequential, as it only applies to the last phase of unspooling when the skein is almost completely unravelled.

    Figure 6 shows a numerical solution of (4.14) for our reference parameter values and with a strain rate λ = 10 s−1 for ℘ = 10. (We choose λ such that λLmax is of the same order of magnitude as U = 1 m s−1 in the pinned cases.) The lower bound (4.17) on the depletion time is 1.48 s, and the numerical value is tdep1.73 s. This is slower than what we observed in the two pinned cases (tdep<1 s), due to the factor 2log(Lmax/L0)14.8. The slowdown due to the short initial thread length is thus considerable in this case. A longer initial length or a higher strain rate would be needed to make the times comparable.

    Figure 6.
    Figure 6. 

    4.4. Two free skeins (skein splitting)

    Another scenario of unravelling is when a skein splits into smaller connected fractions, which then unravel. This scenario is possible since the hagfish ejects the skein through its slime gland and the resulting shear forces (from the ejection process and the fluid’s viscous drag) can break the skein into two.

    Here we consider the simple case of a skein breaking into two halves. The unravelling may be faster since the initial viscous drag is dominated by two skeins, rather than a skein and a small initial length of thread. A diagram of this configuration is show in figure 7: we model the broken skein as two spheres, of radius R1 and R2, respectively, connected by an unravelled length of thread, which can unspool at both ends. We fix a reference point s = 0 between the two skeins such that x(0,t)=X(t). The thread then extends a length L1(t) towards the first skein (right) and L2(t) towards the second skein (left), with L=L1+L2 the total unravelled length. Without loss of generality we take L1(0)=L2(0)=L0/2. The peeling force at the first skein (s=L1x=X+L1) is the sum of the drag forces due to the second skein (s=L2x=XL2) and drag on the thread,

    T|s=L1=6πμR2(u(XL2,t)(X˙L˙2))4πμLδ(u¯(L,X,t)X˙),
    4.19
    where now u¯(L,X,t):=(1/(L1+L2))XL2X+L1u(x,t)dx. Since the thread and skeins are free, the peeling force at s = L1 (4.19) must balance the viscous drag force on the first skein,
    T|s=L1=6πμR1(u(X+L1,t)(X˙+L˙1)).
    4.20
    Equating (4.19) and (4.20), we can solve for X˙,
    X˙=3R1(u(X+L1,t)L˙1)+3R2(u(XL2,t)+L˙2)+2Lδu¯(L,X,t)3(R1+R2)+2Lδ.
    4.21
    We use this to eliminate X˙ from (4.20),
    T|s=L1=6πμR13(R1+R2)+2Lδ(3R2(u(X+L1,t)u(XL2,t)L˙)  +2Lδ(u(X+L1,t)u¯(L,X,t)L˙1)).
    4.22
    We can then also carry out the same calculation for the second skein, at s=L2, and find
    T|s=L2=6πμR23(R1+R2)+2Lδ(3R1(u(X+L1,t)u(XL2,t)L˙)2Lδ(u(X+L2,t)u¯(L,X,t)+L˙2)).
    4.23
    Now it is a matter of solving the coupled peeling equation α(L˙1)m=T|s=L1α(L˙2)m=T|s=L2. To keep things simple, let us take a symmetric configuration centred on x = X = 0 where the two skeins are initially of equal size. (Unequal splitting would result in a depletion time in between this case of even splitting and the free skein–thread of §4.3.) We take an antisymmetric velocity field u(x,t)=u(x,t) that pulls apart the skeins, such as for an extensional flow u=λx. Then R1 = R2 and L1=L2 for all time, and ū = 0. The tensions (4.22) and (4.23) are then equal and greatly simplify to T|s=L1=T|s=L2=6πμR1(u(L1,t)L˙1). Thus, the dynamics for this case is governed by
    α(L˙1)m=6πμR1(u(L1,t)L˙1).
    4.24
    The drag force on the thread has dropped out, since the antisymmetric velocity field leads to cancelling forces on the thread. Another way to think of (4.24) is to observe that in making a symmetric configuration, with the two skeins being pulled apart by a straining flow centred on the origin, we have effectively ‘pinned’ the thread at x = 0. We have thus recovered our pinned thread equation (4.2) from §4.1, with the notable difference that now we cannot use a constant velocity field U, but must resort to a straining flow λx or some other non-uniform flow.
    Figure 7.
    Figure 7. 

    We non-dimensionalize (4.24) using a characteristic length scale R0 and obtain

    (L˙1)m=R1(u(L1,t)L˙1),
    4.25
    where L˙=L˙/λR0 and R=R/R0 are the non-dimensional unravelling rate and skein radius, respectively. The natural dimensionless number in this case is
    =6πμλR02α(λR0)m=6πμR02mλ1mα1.
    4.26

    Figure 8 shows the unravelling dynamics associated with a split skein using parameter values similar to the free configuration of §4.3 and figure 6. The free skein unravels faster when split, as expected (0.756 s versus 1.73 s), owing to a stronger effective drag force and a kinematic upper bound with a rate λ rather than λ/2. There is an important difference between using the free thread–skein equation (4.13) and the free split-skein equation (4.24): the former has a drag slaved to a short initial thread length, whereas for the latter the drag depends on the initial radius of the split skein, which can easily be larger.

    Figure 8.
    Figure 8. 

    In addition to the four cases discussed in this section, we also analysed a slightly more realistic scenario of suction flow where velocity decays away from the mouth of the predator and we consider a pinned skein at different locations away from the mouth. We use an approximate flow profile from the experimental data available in the literature (electronic supplementary material, section III). In general, such a flow profile is both spatially and temporally varying, but we neglect time-dependent variations for our analysis. The peak velocity (at the mouth of the predator) was chosen to match the characteristic velocity (U = 1 m s−1). The velocity decays over a characteristic length scale of the order of the gape size (e.g. opening size of the mouth). We estimate gape size from the video evidence [8] of slime deployment, resulting in extensional strain rates between 0.28–2.2 s−1, with the rate being highest at the predator's mouth. These are smaller extension rates than considered in the earlier cases of this section. The choice of pinning location drastically affects the unravelling time. A depletion time of ∼0.4 s was obtained for the case where the skein is pinned at a distance equal to one-third of the gape size of the predator. This is longer than the unravel time we found for the pinned skein in a uniform flow (≈0.22 s), due to the decaying velocity away from the predator’s mouth, but the unravelling time still falls close to natural unravelling time scales. More complicated spatially and temporally varying flow fields can be treated in a similar way; we expect the time scales in such cases to be of the order of those we found, given that in real scenarios the thread–skein system can be very close to or in the mouth of the predator, and the non-dimensional quantity  is likely to be sufficiently large.

    5. Discussion

    5.1. The role of the dimensionless parameters  and m

    The unravelling times for various cases discussed in §4 depend on the dimensionless quantity , but also separately on the model parameter m in the peeling force law. This is clear from the dimensionless governing equations in §4 that depend on these two dimensionless parameters separately, although m also appears in the definition of . The power-law exponent m determines the peeling force dependence on the unravelling rate. Such a rate dependence exists in peeling scenarios due to the viscoelastic nature of adhesion at the peeling site. In the case of hagfish thread peeling from the skein, the dependence can possibly arise from viscoelastic time scales involved in the deformation of mucous vesicles or the polymeric solution of mucus [15], or the protein adhesive between the loops of thread [11]. The peeling resistance also depends on the dimensional constant factor, α, but its influence on unravelling is built into the dimensionless factor , for which ℘ ∼ α−1.

    Figure 9 compares all four flow scenarios of §4 as a function of  and m in terms of the effective deployment time, i.e. the time to unravel half of the thread length, tdep,50%. (This effective time is used because some flows cannot fully deplete the skein in finite time for m < 1/3, as discussed in §4. Moreover, in practice, the threads do not need to be fully unravelled to create slime.) The flow parameters are identical to those previously described. For all cases, the limit of high drag and low peel resistance, ℘ ≫ 1, converges to the kinematic limit of unravelling where unconstrained portions of the skein–thread system exactly advect with the local flow velocity. At the other extreme, viscous drag is weak compared with peel resistance and for some small value of  unravelling is too slow to match physiological time scales.

    Figure 9.
    Figure 9. 

    The power-law exponent m is a secondary effect compared with . In general, for ℘ > 10, m has negligible effect on unravelling times. For ℘ < 10, the dependence on m is case specific. For the cases of pinned thread (§4.1), pinned skein (§4.2) and free skein–thread (§4.3), a larger value of m leads to a smaller unravelling time (while keeping the same value of ). For the case of skein splitting in an extensional flow (§4.4), such a monotonic trend is not observed and above a critical value of  the unravelling is faster for small values of m. This presumably arises from the nonlinearity in the peel force constitutive equation. For example, taking m = 1 as a reference case, making m < 1 increases the dimensionless peel resistance for L˙<1, but decreases the peel resistance for L˙>1. As such, whether the unravelling rate is L˙1, the exponent m can accelerate or decelerate the unravelling process.

    The value of m affects the minimum required ℘min to achieve unravel times comparable to physiological time scales (i.e. tdep,50% at or below the dotted lines in figure 9). For the uniform velocity field cases (U = 1 m s−1), ℘min is a weaker function of m for the pinned thread case, ℘min = 0.29–1.32, compared with the pinned skein case, ℘min = 0.03–3. For the cases of a free skein–thread in extensional flow, even with splitting, tdep,50% never falls below 400 ms even at high . That is, tdep,50% is higher than the physiological unravel time scales by a factor of 2 or 3. However, as stated earlier, the time scales in such cases are determined by the specific choice of strain rate, λ, and the initial unravel length L0. Such kinematic and geometric parameters are certainly variable in reality, and small changes could easily decrease the unravel time scales, as previously discussed. In any case, m becomes important only when depletion time scales are much larger than the kinematic limit, clearly showing that m is of secondary concern compared with .

    An important caveat to the m = 0 case of FP = α = constant is that peeling cannot occur (L˙=0) if the viscous drag falls below a critical value. For example, if either the initial skein radius R0 or thread length L0 is too small, the viscous drag force is less than FP and unravelling cannot occur. Thus, in figure 9 a minimum value of  is needed for the m = 0 cases. The minimum values range from about 1 to 10, depending on the case and the corresponding definition of  for the flow and geometry. In three cases, the viscous drag can potentially increase during unravelling as the thread elongates (figure 9bd). In these cases, the minimum  is associated with initiating the peeling process. For the other case of the pinned thread, figure 9a, the viscous drag decreases during unravelling, since it is slaved to the skein radius, which decreases in size during the process. Unravelling here will eventually stop at a critical value of R. This, therefore, feeds back to requiring a larger critical initial value of  to unravel by 50%, and is used in figure 9a to determine the domain of  for the m = 0 case.

    5.2. Estimating the parameter 

    A key question remains: what is  in physiological scenarios? For this, we must know the peeling force parameters in the constitutive model, and no direct experimental measurements are yet available. Here we make estimates for the two extreme conditions of m = 1 and m = 0, i.e. a linear dependence on velocity (akin to a constant viscous damping coefficient) and a constant peel force, respectively.

    For the m = 1 case, we consider viscous resistance acting at the peel site with stress σ=μuϵ˙, where μu is the uniaxial extensional viscosity between the separating thread and the skein (related to shear viscosity as μu = 3μ), and ϵ˙=L˙/Lc is the local extensional strain rate that depends on the peel velocity L˙ and the characteristic velocity gradient length Lc. The stress acts over the characteristic thread–thread contact area, which we assume scales as A ≈ d2, i.e. contact across the diameter and the length of contact along the thread also scales with the diameter. The peel force is then

    FPμu(L˙Lc)d2.
    5.1
    Comparing with the peeling law in (3.15), FP=αL˙m, we get
    α=μud2Lc;m=1.
    5.2
    Substituting this into ℘ = FD/FP, and considering the majority of cases where drag is set by the skein radius R0, i.e. equations (4.4), (4.16), (4.26), we obtain
    =6πμR0Uμu(L˙/Lc)d2=6πμμuUL˙R0Lcd2,
    5.3
    where important ratios have been grouped. The simplest case is peel viscosity arising from the surrounding viscous liquid at the peel site. In other words, the viscosity causing drag is also resisting peeling, and to cast in terms of extensional viscosity we take μu = 3μ, a result for a Newtonian fluid. The viscosity μ may be that of sea water, or a surrounding mucous vesicle solution with higher viscosity, but under these assumptions the ratio μ/μu is still the same. Furthermore, typically U/L˙1, and the velocity gradient length scale is likely set by the thread radius, Lc ≈ r. Then (5.3) dramatically simplifies to
    =2πR0d,
    5.4
    which clearly estimates ℘ ≫ 1, or more specifically for R0 = 50 μm and d = 2 μm, ℘ ≈ 160. If drag is instead dominated by the thread length L0, e.g. for the pinned skein case of §4.2, then the numerator in (5.3) would be modified by replacing 6πR0 with 4πL0. We expect L0 to be of the same order as R0, e.g. a single curl in the coil. But if L0 is smaller, it would decrease  accordingly.

    Our specific assumptions can modify the details, but in general we estimate that physiological conditions for m = 1 would give ℘ > 1, if not ℘ ≫ 1. The velocity gradient length scale Lc could be smaller than the thread radius r. A decrease in Lc/r makes  proportionally smaller, but it is difficult to imagine this being more dramatic than, say, a factor of 10. The velocity ratio U/L˙, if anything, will be larger than 1, and this proportionally increases the estimate of . The viscosity ratio μ/μu could be smaller, e.g. if the viscosity of proteins between thread wrappings is larger than the surrounding viscous liquid. However, we note that the surrounding viscous liquid can have a very large viscosity, e.g. the measured extensional viscosity of hagfish mucous vesicle solutions obtained by Böni et al. [15] is μu ≈ 10 Pa s. This is much higher than water, μu ≈ 3 mPa s. An additional mechanism of increasing drag, and , is for mucous vesicles to bind on the thread during the unravelling process (figure 10), which would transmit additional forces to the drag term, as suggested by Winegard & Fudge [22]. Such a scenario is possible since mucous vesicles and thread cells are densely packed inside the slime glands and are released simultaneously. For all of these variations, ℘ > 1 seems very likely for physiological conditions in this constant viscosity estimate for m = 1.

    Figure 10.
    Figure 10. 

    To estimate physiological  for m = 0, the other extreme of a constant force resisting peel, we consider peel strength interactions between the skein fibres solely due to van der Waals forces. We estimate FP = α ≈ 10−7 N (see electronic supplementary material, section II). Substituting into ℘ = FD/FP, and considering cases where drag is set by the skein radius, i.e. equations (4.4), (4.16), (4.26), with R0 = 50 μm, water viscosity μ = 1 mPa s, and U = 1 m s−1, gives ℘ ≈ 90. We see that ℘ ≫ 1 with these assumptions. Even if the force resisting peeling is larger by a factor of 10 or 100, still 1 and viscous hydrodynamics can provide rapid unravelling that can be very close to the kinematically derived lower bounds on unravelling time.

    6. Conclusion

    Our analysis shows that, under reasonable physiological conditions, unravelling due to viscous drag can occur within a few hundred milliseconds and is accelerated if the skein is pinned at a surface, such as the mouth of a predator. A dimensionless ratio of viscous drag to peeling resistance, ℘ = FD/FP, appears in the dynamical equations and is the primary factor determining unravelling time scales. Large  corresponds to fast unravelling that approaches a kinematic limit wherein free portions of the thread–skein system directly advect with the local flow velocity. For characteristic velocity U, the bound is tdepLmax/U, whereas for extensional flows with strain rate λtdepλ1log(Lmax/L0), where L0 is the initial thread length.

    The modelling approach captures essential features and insights by considering a single skein unravelling in idealized flow fields. Future modelling efforts could build on our work by expanding and detailing several aspects, primarily with new experimental measurements of peel strength for skein unravelling, but also details of physiological flow fields including characteristic velocities and strain rates. Real physiological scenarios are more complex due to chaotic flows and multi-body interactions (multiple skeins, mucous vesicles). Our model does not consider such interactions, or the important feature of unravelled threads interacting to create a network. At leading order, we expect such modelling to require more complex flow fields that create extension (to unravel fibres) but also bring different fibres together. Mixing flows would be excellent candidates for theoretical analysis, and any experimental characterization of physiological flow fields should keep this perspective in mind, e.g. simple suction flow with extension, but no mixing, may not be sufficient to create a network of unravelled threads.

    Although the physiological flow fields may be different from the ones that were used in the analysis, our results underline the importance of viscous hydrodynamics and boundary conditions on the process. Recent work [11] found that Pacific hagfish skeins undergo spontaneous unravelling in salt solution. However, the observed unravelling time scales (∼min) are much larger than the physiological time scales (approx. 0.4 s) during the attack. It is possible that ion transport to the peeling site may help in peeling the adhesive contacts, which may be diffusion limited without flow. Although it is known that flow is required to accelerate unravelling to tdep,50%<1 s, it is not yet clear whether flow-enhanced ion transport may also contribute to a faster unravelling, in addition to the drag effects. The effects of various salt ions on the swelling and rupture of mucin vesciles have been studied in the past [18,19], but the influence of such ionic effects on the skeins and their deployment is not yet known. If ion transport and chemistry are important, this would modify the FP behaviour and require modelling of transport at the peel site due to flow. Our results do not rule out the possibility of ion-mediated unravelling but provide an alternative mechanism of unravelling which may be occurring alone or in conjunction with a multitude of other processes.