October 2020 Wetland Science & Practice

Page 1

Wetland Science Practice

&

published by the Society of Wetland Scientists

Vol. 37, No. 4 October 2020 ISSN: 1943-6254

SWS - Celebrating Our 40th Anniversary in 2020


FROM THE EDITOR’S DESK Greetings to all. As we wrap up our issues for 2020, the corona virus still affects everyday life and fires continue to burn in the western United States and in South America’s Pantanal, claimed to be the world’s largest tropical wetland. These fires are devastating and cause for great concern not only for residents but also for natural resource managers. For perspective, in California alone, an area the size of Rhode Island has burned so far this year. In one week, more land burned in California than in all of 2018 and twice as much as had Ralph Tiner burned in 2017 according to WSP Editor the state’s forestry department. In the Pantanal, the number of fires this year are over 10 times the number that occurred in 2018. While lightning is responsible for the California blazes with a couple of exceptions, the combination of lightning and intentional human-set fires (e.g., to convert forest to cattle ranches) are the causes for those in the Pantanal (see some articles listed in “Wetlands in the News” for details). On September 15 and 16, I even witnessed the effect of the U.S. western wildfires on our skies in Massachusetts at sunset (see images below; photos by Barbara Tiner). In one image you’ll see the sun appears as a red dot! Didn’t smell any smoke as it was thousands of feet above the land surface. Along the U.S. East and Gulf Coasts, hurricane season is in full swing with AccuWeather predictions for as many as 28 names storms and 13 hurricanes. To date we’ve had eight landfalls and we’ve still got a couple of months to go before the season ends (normally we get 3 or 4 landfalls each year). All this plus more news on melting glaciers and permafrost. Climate change is real and affecting both wildlife and human populations. Anything positive? For sports lovers, there is something to watch and for all, there are more opportunities for dining out, especially while the weather is good. Interesting times indeed. Also we’ve got a great issue for you! This issue of Wetland Science & Practice has a theme – Latin American wetlands. Last fall in speaking with Tatiana Lobato-de Magalhães about her article for the January 2020 issue, I mentioned that it would be great if we could have an entire issue emphasizing ongoing wetland work in South and Central America. She agreed and took the lead in contacting potential contributors while I contacted a couple of colleagues who I knew were workFrom the Editor’s Desk, continued on page 223 214 Wetland Science & Practice October 2020

CONTENTS Vol. 37, No. 4 October 2020 ISSN: 1943-6254 214 / From the Editor’s Desk 216 / President’s Message 218 / SWS News 219 / Awards 220 / SWS Events 222 / SWS Webinars 223 / Wetland Practice: Passing of a Leader in Wetland Policy Development 224 / INTRODUCTION TO THIS ISSUE: EMPHASIS ON LATIN AMERICAN WETLANDS 231 / Statement on Pantanal Fires ARTICLES Wetland Research 232 / Wetland Science in Latin America and the Caribbean Region: Insights into the Andean States K.S. Navarro and others 241 / Wetlands of the Coast of Lima: Patterns of Plant Diversity and Challenges for their Conservation H. Aponte 246 / Andes, Bofedales, and the Communities of Huascarán National Park, Peru R.A. Chimner and others 255 / Peatlands of the Central Andes Puna, South America E. Oyague and D.J. Cooper 261 / What is the Flora of the Pantanal Wetland? A. Pott and V.J. Pott 267 / Connectivity of River Floodplains – the Case of Ibera Wetlands after 10,000 Years of Isolation from Parana River J.J. Neiff and others 283 / Urban Wetland Trends in Three Latin American Cities during the Latest Decades (2002-2019): Concón (Chile), Barranquilla (Colombia), and Lima (Peru) C. Rojas and others Wetland Conservation/Education/Outreach 294 / Wetland Conservation Concerns in Southern Mexico T. Lobato de-Magalhães and others 302 / Propagation of Endangered Aquatic Plants: An Experience that Promotes ex situ Conservation and Environmental Education S.N. González Mateos


&

Wetland Science Practice NOTES Wetland Research 308 / Bark Traits: A Predictor for Recognition of Successional Groups in Riparian Forest Species J. Rodrigues da Silva and others 310 / Mangrove Ecological Restoration Inside Pantanos de Centla Biosphere Reserve (Tabasco, Mexico) R.A. Betancourth and others 312 / Diversity of Temporary Ponds from the Guajira, Colombia C.E. Tamaris-Turizo and others 314 / Treatment Wetlands – Experiences in Mexico Armando Rivas Wetland Conservation/Education/Outreach 316 / An Ecosystem-based Approach to Managing Fish, Cattle and Forests on the Amazon Floodplain D.G. McGrath and others 319 / Strengthening Goverance in the Monterrico Multiple Use Natural Reserve: Planning for Conservation with a Bottom-Up Approach A. Silvia Morales 321 / Amphibious Colombia: A Country of Wetlands R. Ayazo Toscano and others 323 / Amphibian Territories in Transition: Socio-ecological Rehabilitation of Wetlands R. Ayazo Toscano and others 325 / Guardians of Wetlands (Los Guardianes de los Humedales): Young Peruvians Committed to Wetlands H. Aponte 326 / We Are Wetlands (@Somos_humedales) K. Paz Arteaga 327 / Urban Wetlands Interactive Platform C. Teutsch Barros 328 / Wetlands in the News 331 / Weltand Bookshelf 333 / What’s New in the SWS Journal - WETLANDS 334 / About WSP/Submission Guidelines

COVER PHOTO: Peatland (bofedal), Ayacucho, Peru. Photo by Héctor Aponte.

Note to Readers: All State-of-the-Science reports are peer reviewed, with anonymity to reviewers.

PRESIDENT / Loretta Battaglia, Ph.D. PRESIDENT-ELECT / Gregory Noe, Ph.D. IMMEDIATE PAST PRESIDENT / Max Finlayson, Ph.D. SECRETARY GENERAL / Leandra Cleveland, PWS TREASURER / Lori Sutter, Ph.D. EXECUTIVE ADMINISTRATOR / Suzanna Hogendorn CONSULTING DIRECTOR / Michelle Czosek, CAE WETLAND SCIENCE & PRACTICE EDITOR / Ralph Tiner, PWS Emeritus CHAPTERS ALASKA / Emily Creely ASIA / Wei-Ta Fang, Ph.D. CANADA / Gordon Goldborough, Ph.D. CENTRAL / Katie Astroth CHINA / Xianguo Lyu EUROPE / Matthew Simpson, PWS INTERNATIONAL / Ian Bredlin, Msc; Pr.Sci.Nat and Tatiana Lobato de Magalhães, Ph.D., PWS MID-ATLANTIC / Jennifer Slacum NEW ENGLAND / Dwight Dunk NORTH CENTRAL / Christina Hargiss, Ph.D. OCEANIA / Phil Papas PACIFIC NORTHWEST / Josh Wozniak ROCKY MOUNTAIN / Ryan Hammons, PWS SOUTH ATLANTIC / Brian Benscoter, Ph.D. SOUTH CENTRAL / Scott Jecker, PWS WESTERN / Richard Beck, PWS, CPESC, CEP SECTIONS BIOGEOCHEMISTRY / Lisa Chambers, Ph.D. EDUCATION / Derek Faust, Ph.D. GLOBAL CHANGE ECOLOGY / Wei Wu, Ph.D. PEATLANDS / Bin Xu, Ph.D. PUBLIC POLICY AND REGULATION / John Lowenthal, PWS RAMSAR / Nicholas Davidson WETLAND RESTORATION / Andy Herb WILDLIFE / Andy Nyman, Ph.D. WOMEN IN WETLANDS / Carrie Reinhardt Adams, Ph.D. STUDENT / David Riera COMMITTEES AWARDS / Siobhan Fennessy, Ph.D. HUMAN DIVERSITY / Kwanza Johnson and Jacoby Carter, Ph.D. MEETINGS / Yvonne Vallette, PWS PUBLICATIONS / Keith Edwards MEMBERSHIP / Leandra Cleveland, PWS WAYS & MEANS / Lori Sutter, Ph.D. SWS WETLANDS OF DISTINCTION / Roy Messaros, Ph.D. Bill Morgante and Jason Smith, PWS REPRESENTATIVES PCP / Scott Jecker, PWS STUDENT / David Riera WETLANDS / Marinus Otte, Ph.D. WETLAND SCIENCE & PRACTICE / Ralph Tiner, PWS Emeritus ASWM / Jill Aspinwall RAMSAR / Nicholas Davidson, Ph.D. AIBS / Dennis Whigham, Ph.D. SOCIETY OF WETLAND SCIENTISTS 1818 Parmenter St., Ste 300, Middleton, WI 53562 (608) 310-7855 www.sws.org Wetland Science & Practice October 2020 215


PRESIDENT’S ADDRESS REFLECTIONS, CONNECTIONS AND THE CHANGING SEASONS I’ve been watching intently and enjoying the seasons change in southern Illinois, from summer to autumn, and thinking about other lovely parts of the world now shifting from winter to spring. With all of the challenges and disappointment 2020 has brought, it has also been a time to slow down and reflect on past travel, fieldtrips, and good times with colleagues from all over the world. The love of wetlands has led many of us to travel far and wide and Loretta Battaglia, Ph.D. also brought us together. Just as our precious waters and wetlands Southern Illinois connect landscapes, so too do they University connect us through shared experiSWS President ences, passion for conservation and restoration, and the desire to discover their intricacies and ecological secrets. These connections run deep. If there has ever been a time to sustain and nurture these connections, it is now. The virtual platforms that have enabled conversations, collaborations and the ongoing business of our Society are admittedly not as much fun as slogging through a swamp or bog together or meeting up in a cozy corner at our annual meeting to brainstorm projects and initiatives. They have kept us going, however, and it is likely they will remain a part of the way we operate long after the Covid-19 pandemic is over. Virtual tools have and will continue to be valuable complements to our conventional mode of operations. They will also help lessen the Carbon footprint of our events while better supporting the needs of a global wetland community that is geographically broad but aims to be inclusive. I hope you will join us in celebrating the 40th anniversary of SWS by participating and tuning in to our first Society-wide virtual meeting (December 1-3, 2020), aptly named Wetland Connections over 40 years. n

216 Wetland Science & Practice October 2020


Wetland Science & Practice October 2020 217


SWS NEWS WORLD’S LEADING AQUATIC SCIENTIFIC SOCIETIES URGENTLY CALL FOR CUTS TO GLOBAL GREENHOUSE GAS EMISSIONS In an unprecedented statement released September 14, 110 aquatic scientific societies, including SWS and representing more than 80,000 scientists across the world, joined forces to sound a climate change alarm. The societies call for drastically curtailed global greenhouse gas emissions to avoid the worst impacts of man-made climate change to fish and aquatic ecosystems. Unless urgent action is taken to reduce emissions, scientists predict catastrophic impacts to commercial, recreational, and subsistence fisheries and human health and global economies. n WETLANDS OF DISTINCTION The newly-designated Wetland of Distinction Quakertown Swamp (Pennsylvania) received some favorable press from the Pennsylvania Intelligencer (https://www.theintell.com/ story/news/local/2020/07/22/quakertown-swamp-honoredas-ldquowetland-of-distinctionrdquo/112701110/), as well as in this news video (https://www.wfmz.com/news/ area/southeastern-pa/quakertown-swamp-gets-wetlandof-distinction-title/article_f6c60f74-e7ea-11ea-838d4707c0188c10.html)! And, the first Wetland of Distinction outside the US has been designated, as featured in Restoring a gem in the Murray-Darling Basin: the success story of the Winton Wetlands (https://theconversation.com/restoring-a-gem-inthe-murray-darling-basin-the-success-story-of-the-wintonwetlands-140337). SWS Past President Max Finlayson co-authored this article about the Winton Wetlands in the Murray-Darling Basin of Australia. In addition to being a story about SWS' first Wetland of Distinction designated outside of the United States, this article details a success story about a community-based wetland restoration venture. Visit https://www.wetlandsofdistinction.org/ for information about this impoartant SWS initiative, and to nominate your favorite wetland to be recognized as a Wetland of Distinction today! n MAKE A DIFFERENCE FOR SWS THIS HOLIDAY SEASON Do your Amazon shopping at smile.amazon.com/ch/481146960 to generate donations for SWS, at no additional cost to you. n

218 Wetland Science & Practice October 2020

PUBLIC POLICY UPDATES County of Maui v. Hawai‘i Wildlife Fund amici brief featured in PBS NOVA podcast SWS' Royal Gardner and his team of attorneys, as well as many of our members and members of our aquatic society partners, played a role in preparation of an amici brief cited earlier this year in the U.S. Supreme Court’s opinion in County of Maui v. Hawai‘i Wildlife Fund. On April 23, 2020, SCOTUS ruled 6-3 that the Clean Water Act (CWA) covers the functional equivalent of direct discharges of pollutants to navigable waters. The amici brief was cited by Justice Breyer, who wrote the opinion for the majority, “Virtually all water, polluted or not, eventually makes its way to navigable water. This is just as true for groundwater. See generally 2 Van Nostrand’s Scientific Encyclopedia 2600 (10th ed. 2008) (defining “Hydrology”). Given the power of modern science, the Ninth Circuit’s limitation, “fairly traceable,” may well allow EPA to assert permitting authority over the release of pollutants that reach navigable waters many years after their release (say, from a well or pipe or compost heap) and in highly diluted forms. See, e.g., Brief for Aquatic Scientists et al. as Amici Curiae 13–28.” Steph Tai, a member of Royal's legal team, was recently interviewed for the PBS NOVA podcast Science in the courtroom: https://www.pbs.org/wgbh/nova/podcast/. The discussion on the Maui case and the brief begins about 14:30. Our work is clearly making a difference! WOTUS Rulings Colorado Some good news: the rule appears to be enjoined in Colorado (where judge did not permit amici briefs): https://www. bloomberglaw.com/public/desktop/document/StateofColoradoThevUSEnvironmentalProtectionAgencyetalDocketNo120/4?1592664339 California The Court denied the motion for a preliminary injunction, and the Navigable Waters Protection Rule will go into effect on Monday. Here is the order: https://sws.org/images/ pdfs/CA_v_Wheeler_PI_decision.pdf n


AWARDS

Introducing the Society of Wetland Scientists 40th Anniversary Award PURPOSE: On the occasion of the 40th Anniversary of the Society of Wetland Scientists the Executive Board would like to recognize high level and sustained contributions to wetland research, practice, education, or communication, or service to the Society. These awards are additional to the awards offered annually by the Society and covers the breadth of professional activity normally undertaken by members. Nominations for the Award are open to all members in good standing through the Chairs of the Chapters, Sections and Standing Committees. PROCESS: 1. Each Chapter, Section and Standing Committee is invited to nominate 1 or 2 of their members (in good standing) for consideration as a recipient of the Society of Wetland Scientists 40th Anniversary Award for high level and sustained contributions to wetland research, practice, education, communication, or support to the Society. This message is being sent to all Chairs and also to all members via the Society’s web page news service. 2. The Chair of the Chapter, Section or Standing Committee, or their nominee, is asked to seek nominations from their members and arrange for the supporting information for their agreed 1 or 2 nominations to be forwarded to the Chair of the Awards Committee by November 13, 2020. In a cover, message the Chair (or nominee) should describe in 3-4 sentences the reasons for the nominations along with a separate statement about how they avoided any potential conflicts of interest in choosing their nominees. The Executive Board will also be invited to make nominations with a particular emphasis on sustained contributions, including for support to the Society through contributions to good governance of the Society. 3. Each nomination should consist of a short (1-2 pages) Curriculum Vitae of the nominee and a succinct 1 page

Student Section Virtual Conference Award Winners The SWS Student Section held a virtual conference in June 2020, and have announced award winners. LIGHTNING TALK WINNERS • First Place: Amanda Loder, Nova Scotia, Canada • Tie for Second Place: Andrea Stumpf, Massachussetts, United States Chelsea Duball, Wyoming, United States • Third Place: Clay Tucker, Louisianna, United States

statement by each of two referees in support of the nomination against one or more of the categories for the award (wetland research, practice, education, communication, or support to the Society). These letters should outline the reasons for the nomination and provide tangible examples or evidence. This information should accompany the nominations forwarded from the Chair of each Chapter, Section or Standing Committee to the Chair of the Selection Committee (fennessym@kenyon.edu) and clearly marked “SWS 40th Anniversary Award Nomination”. 4. In making nominations please consider the outcomes and impact on wetland research, practice, education, communication, or support to the Society, and the period over which these have occurred in addition to any individual or one-off high value contributions. The emphasis is on outcomes and impact, not just outputs (that is, while the quality of research publications is an important indicator, it is not the sole indicator in the assessment of importance and impact). The nomination should summarize the evidence in support of the high level and sustained contributions made by the nominee. 5. The award will comprise a certificate with an inscription reading “Society of Wetland Scientists 40th Anniversary Award for high level and sustained contributions to [wetland research, practice, education, communication or service to the Society]” and be announced at the 2020 Virtual Conference. 6. The decision on the number of awards and awardees will be made by the Selection Committee. Late nominations cannot be accepted and the Committee will not enter into any discussion with other parties about the award processes. The Selection Committee will be headed by the Chair of the Awards Committee, along with the President and two Past Presidents of the Society. n

SWS Grants International Travel Awards The Awards Committee awarded two 2020 International Travel Grants, which will be deferred to 2021. Congratulations: • Rajashree Naik, Rajasthan, India Abstract: “Ecological Status of Largest Saline Wetlandscape of India: A Study of Sambhar Lake under DriverPressure-State-Impact- Response (DPSIR) Framework” • Pankyes Datok, Occitanie, France Abstract: “Investigating the role of the Cuvette Centrale wetlands in the hydrology and Organic carbon fluxes of the Congo River basin” Wetland Science & Practice October 2020 219


SWS EVENTS

2020 Virtual Meeting December 1-3, 2020 The Society of Wetland Scientists (SWS) is excited to announce our inaugural virtual meeting on December 1-3, 2020, themed “Wetland Connections Over 40 Years.” We listened to your feedback and decided that a smaller, virtual meeting is in order to celebrate the SWS 40th anniversary and to showcase our members’ outstanding work in wetlands. Our goal is to offer an abbreviated format for the SWS community to connect, to share research, and to continue expanding our Society’s global network. Be sure to visit our event website for all the details! swsvirtualmeeting.org

There are limited sponsorship opportunities remaining: https://static1.squarespace.com/ static/5967a224725e258a852d731e/t/5f848650b21edc3a8a cd6790/1602520658016/Sponsor+opportunties+SWS+2020updated+10.12.pdf n

Thank you to sponsors:

220 Wetland Science & Practice October 2020


2021 Annual Meeting June 1-4, 2021 The Society of Wetland Scientists is planning the 2021 annual meeting, to occur in Spokane, Washington from June 1-4, 2021. Our theme next year is Wetland Science 2021: Adaptation Drives Innovation. We have selected the very industrious and innovative North American beaver (Castor canadensis) as the conference mascot. Their current role and profound historical influence on watersheds, wetlands and hydrology is a growing field of study and restoration. They also provide a model of wetland management and integration of habitats with relevance to our work. We are developing program symposia now and are compiling local and international research topics that tackle the world’s more challenging wetland management issues.

Although we cannot predict what June 2021 will look like for the ever-evolving world pandemic, we are planning a compelling program of speakers, research and presentations that will have a voice and platform for discussion. If all goes well, we plan on hosting the conference at the Davenport Grand, in the heart of the great town of Spokane. There are extensive parks and open spaces at the doorstep of the hotel, and wildlife refuges and wilderness a short drive away. We are also planning for contingencies, including more, smaller session gatherings and opportunities to remotely share information with folks who may have difficulties traveling. We hope you can join us at the SWS 2021 Annual Meeting! https://www.swsannualmeeting.org/ n

STUDENTS - APPLY FOR SWAMMP PARTICIPATION Deadline: Monday, November 16, 2020 The SWS Multicultural Mentoring Program (SWaMMP) works to increase diversity within the Society of Wetland Scientists and throughout the environmental sciences. SWaMMP enables undergraduate students from underrepresented groups to attend and receive full travel benefits to the SWS Annual Meeting, held in Spokane, Washington, June 1-4, 2021. The Annual Meeting offers students valuable career guidance and opportunities to network with leading wetland science professionals from around the world. SWaMMP provides: • Conference registration, lodging and all travel expenses to the SWS Annual Meeting • Individual career mentoring • Postgraduate and career workshops • Networking opportunities to meet professionals from diverse fields • Opportunity to present a research poster More info: http://sws.org/Awards-and-Grants/sws-undergraduate-mentoring-program-swammp.html n

Wetland Science & Practice October 2020 221


WEBINARS

SOCIETY WETLAND SCIENTISTS

Monthly webinars are offered by the Society of Wetland Scientists (SWS) as a benefit of membership. Once each quarter, in March, July, September and December (marked with an asterisk below), the monthly SWS webinar is open for non-members to attend, at no cost, as well. Spanish language webinars are always free for both members and non-members. ENGLISH: sws.org > Events > Upcoming Webinars

SPANISH: sws.org > Events > Spanish Language Webinars

11/19/2020 | 1:00 pm ET

12/9/2020 | 1:00 pm ET

Resurrecting ‘ghost ponds’ and other approaches in pond restoration and conservation Presenter: Emily Alderton

Urban Wetlands, an opportunity to make sustainable cities * Presenter: Carolina Rojas Quezada

12/17/2020 | 9:00 am ET

3/24/2021 | 1:00 pm ET

SWS History - 40 Years of Globalization * Presenters: Kathy Ewel, Gillian Davies, Fred Ellery, Wai-Ta Fang, Max Finlayson, Luisa Ricaurte

The history of biodiversity origin and humanity future through Cuatro Ciénegas wetlands Presenter: Valeria Souza

2/18/2021 | 1:00 pm ET Assessing Vegetative Species Re-colonization of Commercial Cranberry Bogs Presenter: Kate McPherson

ARCHIVES Did you miss a webinar? All webinars are recorded and archived for complimentary viewing by members on our Past Webinars web page.

Thank you to our 2020 Webinar Series Sponsors For more information on sponsoring the SWS Webinar Series, contact membership@sws.org

WWW.IN-SITU.COM

WWW.FACEBOOK.COM/ WATERRESOURCESHYDROLOGY HYDRAULICSEDUCATION/

#SWSWebinars #SWS

222 Wetland Science & Practice October 2020

WWW.WHITENTONGROUP.COM/

#WetlandScience

HTTPS://WILDNOTEAPP.COM/

#WetlandResearch


IN REMEMBRANCE

Passing of a Leader in Wetland Policy Development Dr. Jon Kusler passed away on October 8, 2020. Jon was a leader in helping guide national wetland and floodplain policy since the 1970s. He wrote hundreds of papers on a variety of topics ranging from wetland regulation, restoration, functional assessment to floodplain management. He was the founder of the Association of State Wetland Managers (ASWM), an organization that will continue Jon's legacy and press for improvements in wetland and floodplain conservation. In 1990 he received the first National Wetlands Awards Lifetime Achievement Award from the Environmental Law Institute in recognition of his dedication to wetlands (http://elinwa.org/awards/recipients/jon-kusler). He was a

mover and a shaker for wetland policy and wetlands now receive more attention in the policy arena in large part due to his efforts and those of his collaborators at ASWM and elsewhere. We'll miss Jon, may he rest in peace. The ASWM has established a memorial page on its website for folks to post sentiments and share stories about Jon (https://www.aswm.org/aswm/10210-jon-kusler-memorial). They have also established the Jon A. Kusler Student Scholarship Award to fund student attendance at ASWM's annual State/Tribal/Federal Coordination meeting (for details, see https://www.aswm.org/aswm/donate). n

From the Editor’s Desk, continued from page 214 ing with South American scientists on high-altitude wetlands. What a fantastic job she did - thanks to Tatiana this issue contains 9 articles and 11 notes on various research, conservation, and public education/outreach efforts in Latin America (LA). Needless to say, this kept me quite busy over the past few months. While the collection is not intended to be a comprehensive overview of the ecology and conservation of LA wetlands, the texts should provide readers with a good introduction to these wetlands and the challenges they face, and what some folks are doing to improve the status of wetlands across the region. I again want to especially thank Tatiana and all of the contributors for their efforts in making this issue possible. Sunset, Leverett, MA on September 15, 2020.

I hope that we can get leaders of various SWS sections and chapters, or others to help coordinate contributions so that we can produce other thematic issues of WSP. If interested in taking on the challenge, please contact me at ralphtiner83@gmail.com. The Wetlands of Distinction team has plans to provide profiles of individual Wetlands of Distinction for publication in WSP on a quarterly basis. Meanwhile, I hope all is well and that a vaccine and effective vaccination program will arrive in 2021 well before our next annual meeting in Spokane, Washington. Happy Swamping and Stay Safe. n

Sunset, Leverett, MA on September 16, 2020.

Wetland Science & Practice October 2020 223


INTRODUCTION

This Issue: Emphasis on Latin America Wetlands Tatiana Lobato de Magalhães, Ph.D., PWS, Co-chair International Chapter of the Society of Wetland Scientists

n behalf of the International Chapter and representing Latin America and the Caribbean, I am proud to announce that we now have the first region-focused Wetland Science and Practice issue - the October 2020 Latin America issue. This SWS effort is an important step to strengthen the Society’s internationalization as well as to spread the word on the Society across the Latin American countries. We invited several wetland scientists and practitioners to publish their research or summaries of their research or public outreach projects. This resulted in publication of nine articles and 11 notes in this issue. Here, we can travel across the huge Latin American region, from Mexico to Patagonia, addressing both coastal and highland wetlands and wetlands in between. This issue provides an introduction to the region’s wetlands and examples of ongoing Latin America wetland science and wetland education/public outreach efforts. Such activities include remarkable wetland biodiversity research in Argentine Parana delta and Brazilian Pantanal; challenges for management and conservation in the Brazilian Amazon, Chile, Colombia, Guatemala, and Mexico; wetland organizations’ database in the Andean States; education and awareness in the Chilean Patagonia and Coastal Peru, among others. Thanks to all contributors and the editor Ralph Tiner for supporting this innovative idea and reviewing/editing the contributions. We hope you enjoy reading about Latin American wetlands and learn something about them and the challenges they face. n

O

PICTORIAL OVERVIEW OF LATIN AMERICAN WETLANDS These images are intended to provide readers with a view of some of the wetlands found across Latin America. While they do not show all the types, they reflect the beauty and diversity of types that occur in this species-rich region. We thank the photographers for their contributions. More images are included in the articles and notes that follow. Estuarine salt marsh (Rocuant-Andalién), Talcahuano, Chile by Christopher Momberg and URBANCOST.

224 Wetland Science & Practice October 2020


Mangroves, Ceará, Brazil by Tatiana Lobato de Magalhães, Ph.D., PWS.

Ephemeral wetlands, Lençóis Maranhenses, Brazil by Tatiana Lobato de Magalhães, Ph.D., PWS.

Wetland Science & Practice October 2020 225


Amazon rainforest (location unknown) by Jonathan Lampel (courtesy of unsplash.com).

226 Wetland Science & Practice October 2020


Freshwater wetland, Los Lagos, Los Rios Region, Chile by Alejandro Jara.

Lacustrine marsh, Panguipulli, Los Rios Region, Chile by Karina Arteaga.

Wetland Science & Practice October 2020 227


Highland emergent wetland (temporarily flooded meadow), Guanajuato, Mexico by Tatiana Lobato de MagalhĂŁes, Ph.D., PWS. (Note: Eleocharis densa is the prominent tall plant in the foreground.)

Cushion plant bofedales, Cordillera Apolobamba, Natural de Manejo Integrado Nacional Apolobamba, and Ula Ula National Fauna reserve (for vicuĂąa), 4695 m elevation, Bolivia by David J. Cooper.

228 Wetland Science & Practice October 2020


Highland wetland (Páramos), El Cajas National Park, Ecuador by Tatiana Lobato de Magalhães, Ph.D., PWS.

Inland salt flat (white strip) along Laguna Colorada, Bolivia’s “Red Lake” by Miguel Navaza (www.flicker.com; Creative Commons CC BY-NC-SA 2.0). Its red color is due to red sediments and algae. This area is part of Eduardo Avaroa Andean Fauna National Reserve and Los Lípez - a Ramsar Wetland of International Importance. Three species of flamingos can be found here.

Wetland Science & Practice October 2020 229


Lago Junin (also called Lago Chinchaycocha), in Peru’s Reserve Nacional de Junin (the largest peatland in the Andes, 45 km long and 16 km wide, covering more than 500 km2) by David J. Cooper.

Sphagnum magellanicum cushions on the edge of a ombrotrophic bog near Ushuaia, Argentina by David J. Cooper.

230 Wetland Science & Practice October 2020


Fen in Colombia’s Park Nacional Sumapaz by David J. Cooper. The large plants are characteristic of the paramos region, and feature species of Puya and Espeletia.

PANTANAL

Fire in Paradise Declaration: A Call for Your Participation Scientists of the world, coming from across all disciplines and working across a wide range of subjects and themes, are very concerned with the extensive fires which have been taking place in the Amazon Rainforests and in the Pantanal Region (shared by Brazil, Bolivia and Paraguay), the planet's largest tropical wetland. In order to express our concern, a document titled "Fire in Paradise: Declaration of World Scientists" has been prepared, asking for a set of actions aimed at addressing the problem. The Declaration can be seen at: https://www.haw-hamburg.de/en/university/newsroom/ news-details/news/news/show/fire-in-paradise-declarationof-world-scientists/

If you agree with it, you can sign the Declaration at: https://forms.gle/r7ArbQvoQYvGk6tk9 and express your opinion about a problem which is taking place in Brazil, but which concerns the whole world. Please feel free to pass this note on to other colleagues. n

Wetland Science & Practice October 2020 231


WETLAND RESEARCH Wetland Science in Latin America and the Caribbean Region: Insights into the Andean States Karol Salazar Navarro1,2, Alexandra Dallely Olortigue Tello2, Héctor Aponte2, and Tatiana Lobato-de Magalhães, Ph.D., PWS 3,4

Abstract: Through a mentoring initiative of the Society of Wetland Scientists International Chapter, including volunteer scientists, students, and early-career professionals, we created a database of 283 organizations focused on wetlands for the Andean States (Argentina, Bolivia, Chile, Colombia, Ecuador, Peru, and Venezuela). This review includes data about organizations spatial distribution among seven countries, their sectors, and their principal research areas as well as Andean Ramsar Sites data. The most representative research areas were hydrogeology, biogeochemistry, and biodiversity (represented by fauna, flora, algae, phytoplankton, and zooplankton). We observed a lack of wetland restoration, mapping, and management research. The academic sector (universities) had the largest number of organizations (176 organizations), followed by government (51), non-profits (46), and the private sector (10). The Andean States harbor a total of 94 Ramsar Sites covering 300,000 km2 distributed in seven countries. Through this review, we highlight the magnitude of wetland science in the Andean States and hope to support a future wetland scientist network for the Latin American and Caribbean (LAC) region and allow LAC scientists to connect internationally. INTRODUCTION The Latin American and Caribbean (LAC) region contains 46 countries and non-independent territories (e.g., Aruba, Curaçao, French Guiana, and Puerto Rico) (FAO 2020). Overall, LAC countries harbor a high number of species of fauna and flora, as well as unique neotropical ecosystems. Countries such as Brazil, Colombia, Costa Rica, Ecuador, Mexico, Peru, and Venezuela are considered “megadiverse countries” because they are home to a large part of the planet’s biodiversity (Cancun Declaration 2002). Additionally, the Neotropics support one of the greatest aquatic plant biodiversity and endemism in the world (Murphy et al. 2019). The LAC region contains important and threatened ecosystems known as “biodiversity hotspots” including the Caribbean Islands, Atlantic Forest (Bra1 Universidad Nacional Tecnológica de Lima Sur, Peru. 2 Universidad Científica del Sur, Peru. 3 ECOSUR (El Colegio de la Frontera Sur), Mexico. 4 Correspondence author contact: tatilobato@gmail.com.

232 Wetland Science & Practice October 2020

zil), Mesoamerica, and the tropical Andes, which require greater effort for their conservation since endemic species of plants and vertebrates are being affected by habitat loss and fragmentation (Myers et al. 2000). Across the extensive LAC region, 350 Ramsar Sites are protecting 700,000 km2 of wetlands and a variety of wetland ecosystems such as mangrove forests, swamp forests with palms, flooded savannah and forests, marshes, and peatlands (Junk and Piedade 2011; Ramsar 2020). Despite the high ecosystem services and economic values estimated for wetlands, especially for mangroves and peatlands, LAC wetlands are extremely threatened and at risk of disappearing. In the last four decades, the estimated wetland losses are around 60% in LAC and worldwide (Junk 2013; Davidson 2014; Darrah et al. 2019; Davidson et al. 2019). One of the aims of the International Chapter of the Society of Wetland Scientists (SWS) is to increase wetland science and conservation in the LAC region. Through an integrated database for LAC, we seek to link scientists and wetland professionals at an international level, as well as to engage young people in wetland activities. To achieve this objective, the SWS International Chapter in collaboration with volunteers of Universidad Científica del Sur in Lima, Peru, including students and early-career professionals, created a database of organizations focused on wetlands for the Andean States. Here we present the results for the Andean States, a region that encompasses seven countries (Argentina, Bolivia, Chile, Colombia, Ecuador, Peru, and Venezuela) with unique biodiversity associated with the Andes Mountains and its varied topography. Our main objectives were to answer these four questions: 1) how many organizations are focused on wetlands?, 2) how are they spatially distributed among the countries?, 3) what are the principal wetland research areas?, and 4) how many Ramsar Sites were designated and how they are spatially distributed? Through this review, we highlight the magnitude of wetland science in the Andean States and offer support for a future wetland scientist network for LAC that allows exchange and connection of Latin American and Caribbean scientists regionally and internationally.


FIGURE 1. Andean States wetlands: a) Coastal wetland, Albufera de Medio Mundo, Peru, b) Flamingos in a mangrove, San Pedro de Vice, Peru, c) Páramo La Culata, Sierra Nevada, Mérida, Venezuela, d) Bofedal Guitarrachayoc, Peru, e) Laguna Colorada, Ramsar Site, Bolivia. Photo credits: (a, b, d) Hector Aponte; (c, e) Tatiana Lobato de Magalhães, Ph.D., PWS.

Wetland Science & Practice October 2020 233


FIGURE 2. Heat maps of wetland organizations distribution in the Andean States and distribution of the major Science fields studied in this region.

FIGURE 3. Principal sectors of the Andean States Wetland Organizations.

234 Wetland Science & Practice October 2020

BACKGROUND The Andean States The Andean States of Argentina, Bolivia, Chile, Colombia, Ecuador, Peru, and Venezuela have a variety of ecosystems and climates from 0 to 6,962 m a.s.l., including dry deserts, tropical rainforests, and the snowy highlands (Guerrero et al. 2011). The varied hydrogeomorphology across this region with the different conditions of humidity, precipitation, and temperature determines the characteristics of diverse wetland types. Overall, there are five major Cowardin system types of wetlands in these countries: marine (coastal wetlands, including lagoons, rocky shores, and coral reefs; Figure 1a), estuarine (deltas, tidal marshes, and mangroves; Figure 1b), palustrine (swamps, marshes, and bogs), lacustrine, and riverine types, plus human-made wetlands. In the palustrine type there are many highlands wetlands (e.g., påramos and bofedales; Figure 1c, d), that have been strongly affected by climate change (Junk 2013). Data Search and Classification To identify wetland organizations, we reviewed proceedings of wetland scientific meetings such as congresses, workshops, and symposia held in countries of the Andean region from 2010 to 2020. Due to the lack of wetland-specific scientific events in many of these countries, we considered meetings of closely related science fields (e.g., limnology, ecology, and botany). We followed these four steps: Step 1: Identification of scientific events held between 2010-2020 to search for organizations focused on wetlands, except for Ecuador, which we included a 2001 wetland event (Table 1). Step 2: Detailed analysis of the title, abstracts, and keywords of each article found in the proceedings of the scientific events. This step allowed us to filter out papers on wetland issues and the respective authors’ affiliated organizations. Step 3: Production of a list of organizations that conduct research on wetlands for each country. Step 4: Classification of organizations through an online search, considering the following aspects: (a) general: name, acronym, and sector of the organization (i.e., private, governmental, non-profit, and academic); (b) research areas: identified major research areas through review of the website of each organization and scien-


tific event publications keywords; (c) location: country, state, city, geographical coordinates of each organization; and (d) point of contact: website link, name of an associated researcher, and email address. In addition, we listed the Ramsar Sites for each country, including site code and name, total area, and geographical coordinates for each country (Ramsar 2020). Data Analysis and Mapping We created descriptive graphs to represent total numbers of organizations per sector, per each country’s regions (organizations located in the country capital and country interior cities), and Ramsar sites for each country, using GraphPad Prism v. 8.4.2. We created heat maps (density maps) to represent the spatial distribution of the organizations, major research areas, and Ramsar Sites in the Andean States, using the Kernel density analysis tool of the ArcGIS v. 9.3. To highlight the wetland research areas developed by each country we created “word cloud” graphs with the keywords of all publications selected in Step 2 of the data search, using an online free-tool (wordclouds.com). This tool works with a list of words, which are displayed in different dimensions and randomly distributed, according to the frequency that each word is found in the list in question.

In some countries, we observed a higher number of organizations in the country’s capital region than in the interior (Figure 4), especially in Argentina and Peru. In the latter country, the difference between the capital (57 organizations) and the interior of the country (16) is most evident. On the other hand, we observed the opposite in Bolivia, Chile, Colombia, and Venezuela. Bolivia did not even present any organization located in the constitutional capital (City of Sucre). The highest number of organizations in Bolivia was found in La Paz (13), which is the main political center of the country. The largest number of organizations in Colombia was located in the regions of Cundinamarca and Antioquia, particularly in cities of Bogota (Colombia’s capital with 13 organizations) and Medellín (11), which are the main political and economic centers of the country. It is worth noting that a similar case occurred in Ecuador where 14 organizations were observed in the capital Quito and eight in the city of Guayaquil, located in the Guayas State, Southern Ecuador. Principal Wetland Science Fields We identified 13 major wetland research areas developed by the 283 organizations (Table 2). The most representative ones in the heat maps were hydrogeology, biogeochemistry, and biodiversity (represented by fauna, flora, algae, phytoplankton, and zooplankton). We observed a lack of wetland restoration, mapping, and management (Figure 2). Overall, the word cloud graphs highlighted keywords as ‘macroinvertebrates’, ‘bofedales’, ‘rivers’, ‘biodiversity’, ‘conservation’, and ‘mangroves’ (Figure 5a). The hydrology field is highly studied in South American countries (Kandus et al. 2018)

FINDINGS Distribution of Wetland Organizations in the Andean States We identified 283 wetland organizations in the Andean States, with the most located in Peru (26%) followed by Colombia (17%), Chile (16%), Argentina (12%), Bolivia (10%), Venezuela (10%), and Ecuador (9%) (Figure 2). The academic sector (universities) had the largest number of organizations (176 organizations), FIGURE 4. Wetland organizations location in the country’s capital and interior of the Andean States. followed by government (51), non-profits (46), and private sector (10) (Figure 3). The university sector encompasses research groups, laboratories, institutions, museums, faculties, graduate programs, research centers, and consequently are more representative in the number of studies presented in the scientific events we analyzed in this review. Usually, in developing countries the academic sector represents a higher contribution to scientific research than other sectors (Kumar 2017). Some of the more important Andean universities include Pontificia Universidad Católica del Perú, Universidad San Francisco de Quito, Universidad Central de Venezuela, Universidad Nacional de la Plata, Universidad de Los Andes, and Pontificia Universidad Católica de Chile (World University Ranking 2020).

Wetland Science & Practice October 2020 235


TABLE 1. Proceedings of scientific meetings of the Andean States.

Country

Scientific meeting

Argentina

VI Congreso Argentino de Limnología CAL, 2014

Bolivia Chile Colombia Ecuador

Peru Venezuela

Source

http://sedici.unlp.edu.ar/handle/10915/69018 https://www.researchgate.net/publication/262876439_MemoIV Congreso Boliviano de Ecología, 2014 rias_del_IV_Congreso_Boliviano_de_Ecologia , III Congreso Boliviano de Botánica, 2016 https://issuu.com/iasa-usfxch/docs/memoria_congreso_botanica , http://sociedadchilenadelimnologia.cl/libro-de-resumenes-xivXIV Congreso Sociedad Chilena de Limnología, 2017 congreso-sociedad-chilena-de-limnologia/ IX Seminario Colombiano de Limnología, 2012 https://revistas.udea.edu.co/index.php/actbio/article/view/331656 Humedales Marino - Costeros Continentales, 2001 https://biblio.flacsoandes.edu.ec/libros/digital/56577.pdf , Segunda Reunión de la Iniciativa Regional para la Con- https://www.ramsar.org/sites/default/files/documents/library/ servación y Uso Racional de Manglares y Corales, 2011 informemanglarescoralesguayaquil2011.pdf https://copehu2017.wixsite.com/copehu2017/libro-de-re I Congreso Peruano de Humedales, 2017, sumenes ; https://es.scribd.com/document/429166928/LIBROII Congreso Peruano de Humedales, 2019 RESUMENES-COPEHU-2019-1-pdf https://revistamipensamiento.files.wordpress.com/2015/10/ X Congreso Venezolano de Ecología, 2012 libro-de-resumenes-x-cve-final.pdf

TABLE 2. Principal research areas of the Andean States’ organizations.

Field of Science

Sub-topics

Wetland wise-use and services

Sustainable aquaculture, Fishery, Economic valuation, Ecological valuation, Ecosystem processes, Carbon stock, Management of wetlands, Sustainable development, Ecotourism.

Focus on ecosystem

Mangroves, Marine and coastal ecosystems, Estuaries, Coastal wetlands, Highland Andean bogs, “Bofedales”, Peatlands, Rivers, Basin, Brackish lagoons, Bolivian Amazonia, Pantanal.

Hydrogeology

Sediments, Groundwater, Geomorphology, Geological characterization, Hydraulic parameters, Reservoir, Hydrology, Physicochemical parameters, Hydro systems, Tropical river fluvial, Tributaries, Lake and river water bodies, Fluvial ecosystem dynamics, Water drainage of rains, Watersheds, Hydrobiological resources, Hydrodynamics, Management of water resources, Model ecohydrology of biodiversity, River valleys, Lotic environments, Riversides, Pampas streams, Geological evolution.

Climate Change and Carbon Storage

Climate change, Particulate organic carbon, Adaptation to climate change, Carbon storage, Climate variability.

Conservation and Biodiversity

Biodiversity, Conservation, Endangered species, Ramsar Sites, Taxonomy.

Fauna

Birdlife, Ichthyofauna, Macroinvertebrates, Benthic Macroinvertebrates, Wildlife, Macrobenthos, Entomofauna, Ornithology, Mastozoology, Herpetology, Bivalves.

Botany

Flora, Aquatic plants, Vascular macrophyte index, Vascular flora.

Algae and Zoo-phytoplankton

Zooplankton, Biological communities, Phytoplankton composition, Algal blooms, Microalgae, Taxonomic composition of microalgae, Cyanobacteria, Protozoa, Perifiton, Diatoms.

Biogeochemistry

Water quality, Water pollution, Physicochemical parameters.

Monitoring and Risk analysis

Environmental impact, Indicators of physical habitat, Bioindicators, Water quality, Vulnerability, Water contamination, Biological evaluation of wetlands, Species monitoring, Toxicity, Bioaccumulation, Heavy metals, Ecohydrological indicators, Limnological status index, Microplastics.

Social and Educational

Research dissemination, Environmental education, Socio-economic evaluation, Environmental policy, Traditional management, Sustainable development.

Ecology

Ecological state, Trophic state, Ecosystem processes, Lotic communities, Evapotranspiration, Plant ecology, Phenology, Population dynamics, Eutrophication, Limnology, Migration.

Restoration, Management and Mapping

Reforestation and Restoration, Environmental legislation, Wetland inventory,, Territorial management, Remote sensing, Environmental enforcement and policy.

236 Wetland Science & Practice October 2020


as well as macroinvertebrates, especially in Argentina and Colombia (Ramírez and Gutiérrez 2014). For Argentina, Chile, and Colombia we observed that the principal words of the word cloud graphs were ‘macroinvertebrates’, ‘algae’, ‘zooplankton’, ‘phytoplankton’, and ‘water quality’, which is coherent with the heat maps’ results (Figure 5 b, d, and e). Wetland biodiversity and conservation is probably the most relevant research area for the Andean States since this region harbors ecosystems such as the tropical Andes (an important ‘hotspot’ considered to be a highly diverse ecosystem) and a large area of tropical forests with diverse wetland types (Myers et al. 2000; Junk 2013). Several organizations have been focusing on specific ecosystems such as mangrove forests and coastal wetlands, particularly in Ecuador, Peru, and Venezuela. Despite the widely recognized importance of the coastal wetlands and mangroves for carbon storage (Hamilton et al. 2020) and protection against natural disasters, they remain highly threatened (Davidson 2014; Davidson et al. 2019). For example, Ecuador reported 80% mangrove forest loss in the last decade as a result of aquaculture development (Hamilton et al. 2020). Yet, Venezuela has focused mainly on wetlands associated with the Orinoco River, coastal deltas and lagoons, which have suffered high deterioration, as well as flora studies (Lárez and Prada 2015; Suárez 2016; Marrero and Rodríguez 2017). Bofedales, a highland Andean wetland ecosystem is highlighted by the word cloud graphs in Bolivia and Peru (Figure 5 c, g). This ecosystem is important due to water and human-food resources, biodiversity, and for livestock activities as well as for climate change and carbon accumulation studies (Maldonado 2014). However, the bofedales are one of the most vulnerable ecosystems in the world that are being affected by climate change, as high temperatures will alter peatland distribution and extension (Dangles et al. 2014; Huamán et al. 2020). Peru also stands out in the study of the use and valuation of ecosystem functions, as well as for studies of avifauna, mainly in coastal wetlands that shelter both local and migratory birds (Rivas and Cuellar 2017) (Figures 2 and 5). Ramsar Sites in the Andean States All countries of the Andes are Ramsar Convention signatories. Overall this region harbors a total of 94 Ramsar Sites covering 300,000 km2 distributed in seven countries, Argentina (23 sites), Ecuador (19), Chile (14), Peru (13), Bolivia (11), Colombia (nine), and Venezuela (five) (Figures 6 and 7) (Ramsar 2020). The first Andean Ramsar Site was designated in 1981 (Carlos Anwandter Sanctuary, Chile), while the most recent one was designed in 2018 (Tongoy Bay Coastal Wetland, Chile) (Ramsar 2020).

Ramsar wetlands are mostly distributed in the Argentinean and Ecuadorian territory. Another large portion of these sites are distributed in Northern Peru and Southern Bolivia (Figure 6). The majority of the Andean Ramsar Sites are beautiful and scenic places with tourism importance (e.g., the Laguna Colorada in Southern Bolivia) (Figure 1e). CONCLUSIONS AND INSIGHTS INTO WETLAND SCIENCE IN THE ANDEAN STATES Even though the Andean region is extremely important for wetland biodiversity due to its diversity of ecosystems as well as endemic flora and fauna (Junk and Piedade 2011; Marreno and Rodríguez 2017; Murphy et al. 2019) throughout the region, there is a lack of wetland-focused scientific events, associations, and societies. We can highlight the following wetland scientific local meetings: Argentinean Congress of Limnology, Chilean Society of Limnology Congress, Colombian Seminar of Limnology, Marine Coastal and Continental Wetlands Meeting, Regional Initiative for the Conservation and Wise-use of Mangroves and Corals Meeting, and the Peruvian Congress of Wetlands (Table 1). We carefully analyzed their proceedings, as well as some local scientific events related to wetlands such as the Bolivian Congress of Ecology, Bolivian Congress of Botany, and Venezuelan Congress of Ecology to highlight the efforts of researchers involved with local organizations. The creation of this Andean wetland organizations database with 283 organizations is a first step to recognize this wetland science community and promote a LAC wetland-network. This review allowed us to identify research areas that stand out in the Andean States such as biodiversity (mainly avifauna, macroinvertebrates, botany, phytoplankton, and zooplankton), ecology, hydrogeology, and water quality. As emerging areas that have the potential to be developed in the Andean region, we believe research in education, climate change, sustainable use, and ecosystem services valuation could be expanded. Yet, we consider that restoration, mapping, management, and wetland policy are less represented in the Andean States and extra efforts are needed to develop and research these important issues in the region. Additionally, identifying wetlands research areas and topics in Venezuela was a challenge due to the scarcity of information about wetlands on digital platforms, probably because the political scenario of the last decades, where wetland science and conservation were not a priority (Lobato-de Magalhães et al. 2016). Overall, we observed a high number of wetland organizations and Ramsar Sites in the Andean States, which reinforces the countries’ government commitment to preserving wetland biodiversity and ecosystem services. The observed number of wetland research organizations reflects Wetland Science & Practice October 2020 237


FIGURE 5. Word cloud graphs using the keywords from research presented on Scientific meetings’ proceedings in the Andean States: a) Overall in Andean States, b) Argentina, c) Bolivia, d) Chile, e) Colombia, f) Ecuador, g) Peru, h) Venezuela.

238 Wetland Science & Practice October 2020


the potential of scientific production in the region, the human capital available to generate it, and efforts to bring this wetland scientific-knowledge closer to the population and decision-makers locally. This review is an approach to understanding the potential of the LAC region to address wetland science and contribute to the international scientific wetland community. It is also important that, in the coming years, alliances are sought between research and conservation groups of LAC and other regions (e.g., Global Environmental Facility “GEF Humedales”, Wetlands International, and World Wildlife Fund). A good example of such joint work is what is currently happening with the Coastal Wetlands Initiative (humedalescosteros.org), which seeks to protect the Pacific desert wetlands corridor and to generate synergies between researchers and partners in Chile, Ecuador, and Peru. Finally, the elaboration of this review reflects the role of the SWS in creating opportunities to integrate scientists from diverse countries. More efforts are needed to further wetland science and conservation in the LAC region. However, review paper initiatives focused on wetland organizations and research topics including other LAC regions can create the appropriate scenario that will allow successful coordination of wetland science and conservation in the LAC region. Additionally, through an SWS-LAC network, we could connect wetland scientists and allow them to disseminate their efforts on wetland research to other Latin American and Caribbean scientists and internationally. n

FIGURE 6. Ramsar Sites distribution in the Andean States.

FIGURE 7. Ramsar Sites in the Andean States, number of sites and total area (km2).

ACKNOWLEDGMENTS We thank D. Faust (SWS Education Section) and D. Lobato for reviewing this manuscript. The corresponding author holds a postdoctoral fellowship from ECOSUR (El Colegio de la Frontera Sur) at CONACyT (National Council of Science and Technology) in Mexico. We appreciate the valuable comments of the editor R. Tiner. REFERENCES

Cancun Declaration. 2002. Grupo de Países Megadiversos Afines. Available via http://www.leisa-al.org/web/index.php/ volumen-17-numero-4/2284-declaracion-de-cancun-de-paisesmegadiversos-afines. Accessed June 27, 2020. Dangles, O., R.I. Meneses, and F. Anthelme. 2014. BIOTHAW: Un proyecto multidisciplinario que propone un marco metodológico para el estudio de los bofedales altoandinos en un contexto de cambio climático. Ecología en Bolivia 49(3): 6-13. Darrah, S.E., Y. Shennan-Farpón, J. Loh, N.C. Davidson, C.M. Finlayson, R.C. Gardner, and M.J. Walpole. 2019. Improvements to the Wetland Extent Trends (WET) index as a tool for Wetland Science & Practice October 2020 239


monitoring natural and man-made wetlands. Ecological Indicators 99: 294-298. Doi:10.1016/j.ecolind.2018.12.032 Davidson, N.C. 2014. How much wetland has the world lost? Longterm and recent trends in global wetland area. Marine and Freshwater Research 65(10): 934-941. Doi:10.1071/MF14173 Davidson, N.C., A.A. van Dam, C.M. Finlayson, and R.J. McInnes. 2019. Worth of wetlands: Revised global monetary values of coastal and inland wetland ecosystem services. Marine and Freshwater Research 70(8): 1189-1194. Doi:10.1071/MF18391 FAO. Food and Agriculture Organization of the United Nations. 2020. Latin America and the Caribbean region. Available via http://www.fao. org/3/v8300s/v8300s0o.htm Accessed July 2, 2020. Guerrero, A.L., S.S. Gallucci, and P. Michalijos. 2011. Países Andinos: aportes teóricos para un abordaje integrado desde las perspectivas geográfica y turística. Huellas 15: 121-138. Hamilton, S. 2020. Mangroves and Aquaculture: A five decade remote sensing analysis of Ecuador’s estuarine environments. Doi:10.1007/9783-030-22240-6 Huamán, Y., P. Moreira, R. Espinoza, R. Llanos, J. Apaéstegui, B. Turcq, and B. Willems. 2020. Influencia de los cambios climáticos en la acumulación de carbono en Bofedales Altoandinos durante los últimos 2 500 años. Ecología Aplicada 19(1): 35-41. Doi:10.21704/rea.v19i1.1444 Junk, W.J. 2013. Current state of knowledge regarding South America wetlands and their future under global climate change. Aquatic Sciences 75: 113-131. Doi: 10.1007/s00027-012-0253-8 Junk, W.J. and M.T. Piedade. 2011. An introduction to South American wetland forests: Distribution, definitions and general characterization. In: W.J. Junk, M.T. Piedade, F. Wittmann, J. Schöngart, and P. Parolin (eds.). Amazonian Floodplain Forests: Ecophysiology, Biodiversity and Sustainable Management. Springer Netherlands. pp. 3-25. Doi:10.1007/97890-481-8725-6_1 Kandus, P., P. Gail, N.S. Morandeira, R. Grimson, G. González, E.B González, L. San Martín, and M.P. Gayol. 2018. Remote sensing of wetlands in South America: status and challenges. International Journal of Remote Sensing 39(4): 993-1016. Doi:10.1080/01431161.2017.1395971 Kumar, V. 2017. The role of university research centers in promoting research. Journal the Academy of Marketing Science 45: 453-458. Doi: 10.1007/s11747-016-0496-3

240 Wetland Science & Practice October 2020

Lárez, A., and E. Prada. 2015. Diversidad florística en humedales de las planicies deltáicas del estado Monagas, Venezuela. Acta Biologica Venezuelica 34(1): 61-74. Lobato-de Magalhães, T., C. Varela Romero, and M. Martínez. 2016. Public policies and legislation for wetland conservation in Latin America. Society of Wetland Scientists Annual Meeting Proceedings, Corpus Christi, Texas, USA. Maldonado M.S. 2014. An introduction to the bofedales of the Peruvian High Andes. Mires and Peat 15(4): 1-13. Marrero, C., and D. Rodríguez. 2017. Humedales asociados a sistemas palustres en ámbitos de planicie de los llanos del Orinoco y peniplanicies del Ventuari-Casiquiare, estado Amazonas, Venezuela. BioLlania 15: 616-624. Murphy, K., A. Efremov, T.A. Davidson, E. Molina-Navarro, K. Fidanza, T.C.C. Betiol, ... and M. Kennedy. 2019. World distribution, diversity and endemism of aquatic macrophytes. Aquatic Botany 158: 103-127. Doi:10.1016/j.aquabot.2019.06.006 Myers, N., R.A. Mittermeier, C.G. Mittermeier, G.A.B. da Fonseca, and J. Kent. 2000. Biodiversity hotspots for conservation priorities. Nature 403(6772): 853-858. Doi:10.1038/35002501 Ramírez, A. and P.E. Gutiérrez. 2014. Estudios sobre macroinvertebrados acuáticos en América Latina: avances recientes y direcciones futuras. Revista de Biología Tropical 62(2): 9-20. Ramsar. 2020. Ramsar Sites Information Service. Available via https:// rsis.ramsar.org/ Accessed July 12, 2020. Rivas, R. and I. Cuellar. 2017. Importancia del humedal de Lurín, Lima, Peru. En Libro de resúmenes: I Congreso Peruano de Humedales. p. 78. Available via https://dc280409-c7d4-4755-bfbf-12d39b2d9792.filesusr. com/ugd/ec3c3b_c4ac539b81164b26a4539b5e2c6bd6d7.pdf Accessed June 27, 2020. Suárez, C. 2016. Uso y abuso de las lagunas costeras venezolanas. Revista de Investigación, 40(87): 53-86. Available via https://www.redalyc. org/pdf/3761/376146819005.pdf Accessed 27 Jun 2020 Accessed June 27, 2020. World University Ranking. 2020. Available via https://www.topuniversities.com/university-rankings/world-university-rankings/2020 Accessed June 27, 2020.


WETLAND RESEARCH

Wetlands of the Coast of Lima: Patterns of Plant Diversity and Challenges for their Conservation Héctor Aponte1

ABSTRACT he wetlands of the coast of Lima are ecosystems very close to the city. This makes them particularly important wetlands in composition and function. In recent years, multiple investigations have been carried out on plant diversity of these wetlands. This allows us to know various aspects of their composition and structure. This article compiles the most important information on these ecosystems (mainly at the vegetation level), trying to establish some patterns regarding their diversity at the alpha, beta and gamma levels. As a result, we found that alpha and gamma diversity is intimately related to human activities, which triggers, for example, the presence of invasive species that now represent approximately 50% of the plants. Plant richness is independent of the state of conservation of the areas (having protected areas of high and low diversity compared to the non-protected) and size (the smallest wetland is the one with the highest values of richness per area unit). At the regional level, no patterns of beta diversity have been found, which suggests that we should conserve each wetland along the corridor. Some of the challenges for the conservation of these ecosystems are raised.

T

(2013) for just six coastal wetlands of Lima reported the presence of 123 vascular plant species. When we add birds (Tello and Engblom 2010), mammals (Pacheco et al. 2015), reptiles (Icochea 1998), spiders (Paredes 2010) and protozoa (Guillén et al 2013; Guillén et al. 2015) to the plant species, more than 300 species of organisms occur in Lima’s wetlands. This diversity alone testifies to the importance of these ecosystems as shelters for the diversity of life on the desertlike coast of Peru. How much do we know about the patterns that diversity of species follows in this region? And, if known, can this help us make better conservation decisions? FIGURE 1. Map showing the location of the seven wetlands studied. MM=Albufera de Medio Mundo, CAR=Humedal de Carquín – Hualmay; PAR=Laguna El Paraíso; SR=Humedales de Santa Rosa; VEN=Humedales de Ventanilla; PAN=Pantanos de Villa; PV=Humedales de Puerto Viejo. Scale bar =100km.

INTRODUCTION Peru is rich in wetlands including coastal, Andean, and Amazonian types. The country has more than 12,000 lagoons: 3,896 in the Pacific basin, 7,441 in the Atlantic basin, 841 in the Titicaca basin and 23 in the closed basin of the Huarmicocha System that contribute to this diversity (Ministerio de Agricultura, INRENA 1996). On the coast of Peru, wetlands are part of the migratory route of birds along the Pacific Corridor and as such are especially important for the conservation of biodiversity. To date, 91 coastal wetlands, 56 natural, 11 artificial, and 14 river mouths are known and of these, 25 are in Lima: 16 natural, 4 artificial and 5 river mouths (Pronaturaleza 2010). More than 80 species of vascular plants and 180 species of birds been recorded in these wetlands by preliminary studies (Pronaturaleza 2010). More detailed work of the different biological groups produce more accurate figures of the diversity that is housed in these ecosystems. For example, the work of Aponte and Cano 1 Professional Wetland Scientist. Associate Researcher, Universidad Científica del Sur, Lima- Perú; haponte@cientifica.edu.pe. Wetland Science & Practice October 2020 241


STUDY AREA In the present study, we worked with data from seven wetlands of the central coast of Peru: Wetlands of Puerto Viejo, Pantanos de Villa, Ventanilla Wetlands, Wetlands of Santa Rosa, El Paraíso Lagoon, Carquín-Hualmay Wetland and Albuferas de Medio Mundo. They are located between 10°58`05.15 “S and 12°34`16.77” S, and occur from sea level to altitudes no higher than 25m (Figures 1 and 2). Most of these wetlands are associated with coastal lagoons. While most of the water of these wetlands comes from the Andes, some water enters the subsoil in the highlands and flows into coastal lagoons. In many cases, this water crosses urban, livestock or agricultural areas before entering the subsoil and upon reaching the coast results in euthrophic ecosystems. In some other cases, it enters through temporary connections (directly or by the underground) between the lagoons and the sea. METHODS To answer the questions raised about diversity patterns, several sources of vascular plant richness data were examined: Aponte and Cano (2013) for the wetlands of Paraíso (not protected), Santa Rosa (not protected), Medio Mundo (protected by the regional government) and Puerto Viejo (not protected); Aponte and Ramirez (2014) for the Ventanilla wetlands (protected by the regional government); Ramirez and Cano (2010) for Los Pantanos de Villa (protected, Ramsar site and a National Protected Area) and Aponte and Cano (2018) for Carquín-Hualmay Wetland (not protected). Complementary to the species richness of each wetland (which can be an estimator of gamma diversity) are other measures of diversity on a different scale. For FIGURE 2. The lagoon and wetlands of Paraíso (Huacho, Lima). Wetlands of the coast of Lima are ecosystems formed by the runoff from the water that comes from the Andes frequently interacting with salt water brought in underground from the ocean. The freshwater entering these wetlands often crosses agricultural fields and urban areas.

242 Wetland Science & Practice October 2020

example, the alpha diversity indices (such as Shannon-Wiener Index) indicate diversity at the habitat level, while beta diversity estimators (such as Whittaker and similarity indices) allow us to evaluate the exchange of species between sites (Halffter and Moreno 2005). Available information on alpha diversity (mean Shannon-Wiener value from different habitats from Aponte and Ramírez 2011) and beta diversity (Whittaker and Cody indices from Aponte 2017a, 2017b) for these wetlands was added. To evaluate the correlation between beta diversity and distance between wetlands, the similarity between plant composition in these wetlands was calculated using the Jaccard index. The equations and detailed interpretation of all the mentioned indices can be found in Moreno (2001). The distance between the wetlands was measured using Google Earth. RESULTS AND DISCUSSION Alpha and Gamma Diversity: Their Relationship with Anthropic Processes The data obtained to date show that alpha (habitat level) and gamma (wetland level) diversity is high in both protected wetlands and non-protected wetlands (Table 1). For example, Los Pantanos de Villa (a protected wetland) has the greatest historical richness reported to date while the Santa Rosa wetland (non-protected) is the second-ranked wetland with the highest number of plant species and alpha diversity, despite receiving minimal protection and being located where human activities have considerably intervened the landscape. In the latter case, livestock seems to play an important role in increasing diversity at the alpha level (with high-value plant communities where this activity is practiced). Finding an adequate level of human activity is essential to achieve an adequate use of space and the consequent conservation of the ecosystem. One of the key challenges for the conservation of biodiversity has been the need to conserve the “natural” diversity of ecosystems, so that not only the species found in ecosystems are conserved, but also those that historically have a role in that place (Santana 2019). To achieve this in the wetlands of the Peruvian coast, it is essential to carry out historical studies that allow us to understand the temporal composition of these ecosystems, in order to identify the natural diversity of plants while maintaining the ecosystem processes and services that these species provide. Likewise, it is important to identify the level of contribution that non-native species have in the ecosystem and in the ecosystem services they provide in order to adequately decide what steps to follow in their management. It has been suggested that non-native species should be considered part of the diversity of ecosystems, at least initially (Schlaepfer 2018). So we must consider their distribution in ecosystems such as Peruvian coastal wetlands


where non-native plants represent a considerable percentage of the species in the plant community. The most recent study of plant diversity carried out in the Carquín-Hualmay wetland revealed the importance of small wetlands (which are often little studied) (Aponte and Cano 2018). Although this wetland was not the most diverse of those along the coast of Lima, due to its size, it hosts the greatest number of species per area. This makes it one of the most important ecosystems in the region, and by this alone, worthy of conservation. Moreover, its location at the center of the coast of Lima makes it vital as the connection in the diversity corridor along the coast. Finally, the origin of the flora of these wetlands shows that approximately 50% of the plants are invasive species (see the tables in Aponte and Cano 2013 for more details). This situation is the result of the interaction of these ecosystems with adjacent populations and human activities (e.g., garbage disposal, livestock grazing, and agriculture; Young 1998; Ramirez et al. 2010). The presence of invasive species is now a characteristic of these ecosystems and occurs both in protected and unprotected wetlands. While the presence of these species is typically interpreted as a symbol of the deterioration of wetlands, it might be interpreted it as an integral part of surviving in a humaninfluenced landscape that accomplish the function as a reservoir for urban flora. So our conservation goal must reflect the reality of invasive species being a component of the diversity of these wetlands, but the challenge is how to manage them. Regional Beta Diversity: Patterns in Plants? The studies carried out at the beta level show that, although birds have a greater richness of species in Lima, the turnover between sites (beta diversity) is greater in plants than in birds (Figure 2). This means that the complementarity between wetlands at the regional scale is greater in the case of plants than in birds. Considering this indicator, it is very difficult to decide which wetland to conserve and which one not to conserve, or rather where do we prioritize conservation initiatives? Also, by analyzing the diversity from this second perspective (considering complementarity), we can understand the importance of each of these wetlands for the conservation of the region’s flora. When analyzing the correlation between beta diversity and distance between sites, no pattern of beta diversity at the regional scale was found. Normally, the greater the distance between localities, the greater the difference between them and as a consequence there is a greater beta diversity, but this pattern was not observed in the study area (Figure 3). This indicates that the generalization of known patterns for diversity is not the best route to make conservation decisions. In this case, the structure and composition

of plant species is governed by a set of processes where human activities intervene. These interactions are often complex and little studied. Consequently, it is not possible to easily make a decision on which ecosystem to protect and makes it is necessary to study the processes that occur within and between these wetlands in order to make a good conservation decision. For example, little is known about the role of connectivity in these ecosystems and between these wetlands and coastal hills. Recently the presence of an aquatic plant has been reported in the lomas formations – Lemna minuta Kunth (duckweed). This plant most likely came from a nearby wetland (Aponte 2016). The absence of patterns among these wetlands, as well as their high complementarity, is an indicator that the best way to protect them is through the creation of large-scale protected areas. There is experience of this type in Peru (i.e., the island and guano islands protection system), so these proposals are not unfamiliar to decision-makers. Currently there are initiatives to protect the wetlands of the Pacific coast (for exFIGURE 3. Diversity of birds and plants in the coastal wetlands of Lima, evaluated with the beta indices of a) Whitaker and b) Cody, and c) at the richness level.

FIGURE 4. Scatter plot of the distance between wetlands and the similarity (measured with the Jaccard index) between them. The figure shows no correlation between beta diversity and distance between sites which means that there is no spatial pattern of beta diversity at the regional scale. The correlation analysis for these variables resulted in p> 0.05.

Wetland Science & Practice October 2020 243


ample the Initiative for the Conservation of Coastal Wetlands and Shorebirds in the Arid Coast of the South American Pacific), where these proposals would fit perfectly. Challenges for the Conservation of the Coastal Wetlands of Lima The study of diversity patterns identified research needs that are important to make adequate conservation decisions. There are small wetlands that have not yet been studied (e.g., Laguna la Encantada) and whose importance as shelters for diversity should not be underestimated. Likewise, the role of invasive plant species (considering that, in some cases, they represent around 50% of the species) in these wetlands is hitherto unknown, so it is necessary to study them before indicating that they are harmful species TABLE 1. Plant mean alpha diversity, historical gamma diversity, area and number of species per area for the studied wetlands. Protected wetlands are marked with a *

Wetland Puerto Viejo Pantanos de Villa* Ventanilla* Santa Rosa Carquín-Hualmay Paraíso Medio Mundo*

Mean Gamma Alpha 0.32 32 0.72 1.05 0.48

72 37 67 41 33 21

Area (ha) 200

Species/ area 0.16

276 265 60 11.64 690.42 261

0.26 0.14 1.12 3.52 0.05 0.08

FIGURE 5. A pelican in the coastal wetland Poza de la Arenilla (La Punta, Callao), an artificial wetland that is enriched by nutrients (probably from urban origin plus natural organic matter) causing the growth of thick mats of green algae (in the background). Controlling the growth of these algae and the sources of eutrophication is very important, and will only be achieved with research and appropriate land management. To date, no studies of the flora of this wetland have been published.

for those ecosystems. In addition to these, Lima’s wetlands have several common characteristics (e.g., proximity to the city, land ownership or management problems in perimeter areas, presence of human activities for the use of its resources, absence of permanent monitoring programs for both biotic and abiotic components, and ignorance of the existence and importance of these ecosystems by the general public) that have allowed us to identify situations requiring immediate attention for conservation (Aponte 2017c; Aponte et al. 2018). At least three activities have been identified as fundamental for the protection of these ecosystems in the long term: • Legally recover the spaces. Many of the areas have owners who could legally eliminate the wetland. NGOs and the state must work to reverse the land tenure processes and be able to properly manage these spaces. This will require public education and outreach and dialogue with the owners to help them understand the environmental value of the wetlands and how use of these lands can be done in a way to minimize adverse effects. • Media movement. It is necessary to involve local media (radio, television) and social networks to help improve the public’s understanding and appreciation of these wetlands. Surprisingly, many people who live in Lima do not know about the existence of these ecosystems. This outreach must include information (e.g., pamphlets and videos) and education through field trips by school students and encouraging universities to use wetlands as a subject for ecosystem research. • Working together. It is essential that all people involved in the protection of these wetlands work together. In recent years there have been several personal and isolated initiatives, but I think it is more productive to work cooperatively on outreach activities. Environmental professionals need to work with economists, lawyers, sociologists and communicators on improving public awareness of wetlands and conservation needs. In this way, the fruits of interventions and research can be analyzed from various perspectives, considering possible gaps in the design of management and management plans that allow for solutions (Figure 5). Along with the lomas formations where we also find hundreds of species (Dillon et al. 2011), wetlands are an important refuge for the flora and fauna of this region. Due to their proximity to human populations, both lomas formation and wetland ecosystems have suffered reductions in their extent and degradation by various impacts.

244 Wetland Science & Practice October 2020


Therefore, management plans must be applied in an integral manner for the management of the coastal ecosystems of Lima and Peru. n ACKNOWLEDGMENTS The present study is the result of multiple researches achieved thanks to the support of the Museum of Natural History (Universidad Nacional Mayor de San Marcos), the Universidad Científica del Sur, NGOs such as Terra Nuova, Instituto Peruano por la Sostenibilidad y el Desarrollo and Cooperacción, and the hard work together with colleagues, students and local people in the field, to whom I am very grateful. REFERENCES

Aponte, H. 2017a. Diversidad beta en los humedales costeros de Lima, Perú: estimación con índices de presencia/ausencia y sus implicancias en conservación. Biologist (Lima) 15 (1):9–14. Aponte, H. 2017b. Un ajuste a la diversidad beta en los humedales costeros de Lima. The Biologist (Lima) 15: 479-481 Aponte, H. 2017c. Humedales de la Costa central del Perú: Un diagnóstico de los humedales de Santa Rosa, laguna El Paraíso y Albufera de Medio Mundo. Cooperacción, Lima - Perú Aponte, H. 2016. Nuevo registro de flora para las Lomas de Lachay: Primer reporte de Lemna minuta Kunth (Araceae). Ecología Aplicada 15(1):57–60. Aponte, H. and A. Cano. 2013. Estudio florístico comparativo de seis humedales de la costa central del Perú: Actualización y nuevos retos para su conservación. Revista Latinoamericana de Conservación 3(2):15–27. Aponte, H. and A. Cano. 2018. Flora vascular del Humedal de Carquín - Hualmay, Huaura (Lima, Perú). Ecología Aplicada 17:69–76. doi: 10.21704/rea.v17i1.1175

Guillén, G., E. Morales, and R. Severino. 2013. Adiciones a la fauna de protozoarios de los Pantanos de Villa, Lima, Perú. Revista Peruana de Biología 10:175–182. doi: 10.15381/rpb.v10i2.2500 Halffter, G. and C. Moreno. 2005. Significado biológico de las diversidades alfa, beta y gamma. In: G. Halffter, J. Soberon, P. Koleff, and A. Melic (eds.). Sobre Diversidad Biológica: el Significado de las Diversidades Alfa, Beta y Gamma. m3m-Monografias 3ercer Milenio.vol 4, SEA, CONABIO, Grupo DIVERSITAS & CONACYT, Zaragoza. pp. 5 - 8. Icochea, J. 1998. Lista Roja preliminar de los anfibios y reptiles amenazados del departamento de Lima. Los Pantanos de Villa: Biología y Conservación. Universidad Nacional Mayor de San Marcos, Lima, Perú, pp 217–219 Ministerio de Agricultura, INRENA. 1996. Estrategia Nacional para la Conservación de Humedales en el Perú. Moreno, C.E. 2001. Métodos para medir la biodiversidad. M&T–Manuales y Tesis SEA, Vol. 1. Zaragoza. Pacheco, V.R., A. Zevallos, K. Cervantes, and J. Pacheco. 2015. Mamíferos del Refugio de Vida Silvestre Los pantanos de Villa, Lima, Perú. Cientifica 12(1):26–41. Paredes, W. 2010. Diversidad y variación espacio-temporal de las comunidades de arañas en la Zona Reservada de Pantanos de Villa, Lima, Perú. Tesis para optar por el grado de bachiller en Ciencias Biológicas, Universidad Nacional Mayor de San Marcos Pronaturaleza. 2010. Documento base para la elaboración de una Estrategia De Conservación de Los Humedales de la Costa Peruana. 6–94. Ramirez, D.W., H. Aponte, and A. Cano. 2010. Flora vascular y vegetación del humedal de Santa Rosa (Chancay, Lima). Revista Peruana de Biología 17:105–110. Ramirez, D.W. and A. Cano. 2010. Estado de la diversidad de la flora vascular de los Pantanos de Villa (Lima - Perú). Revista Peruana de Biología 17:111–114.

Aponte, H., D.W Ramírez, and G. Lértora. 2018. Pantanos de Villa: Un oasis de Vida en Lima Metropolitana. Fondo Editorial de la Universidad Científica del Sur., Lima - Perú.

Santana, C. 2019. Natural diversity: how taking the bio- out of biodiversity aligns with conservation priorities. In: E. Casetta, J. Marques da Silva, and D. Vecchi (eds.) From Assessing to Conserving Biodiversity. History, Philosophy and Theory of the Life Sciences. SpringerLink Vol 24: 401-414.

Aponte, H. and D.W. Ramírez. 2014. Riqueza florística y estado de conservación del Área de Conservación Regional Humedales de Ventanilla (Callao, Perú). Biologist (Lima) 12(2):270–282.

Schlaepfer M.A. 2018. Do non-native species contribute to biodiversity? PLoS Biology16(4): e2005568. https://doi.org/10.1371/journal. pbio.2005568

Aponte, H. and D.W. Ramírez. 2011. Los Humedales de La Costa central del Perú: Comunidades Vegetales y Conservación. Revista Ecología Aplicada 10(1):31–39.

Tello, A. and G. Engblom. 2010. Lista de especies de los humedales de la Región Lima: Aves. In: A. Tello and L. Castillo (eds.) Humedales de la Región Lima, Guía de su fauna y flora silvestres. Gobierno Regional de Lima, Lima - Perú. pp. 87–90.

Dillon, M.O., S. Leiva González, M. Zapata, P. Lezama Asencio, and V. Quipuscoa Silvestre. 2011. Floristic checklist of the Peruvian Lomas Formations. Arnaldoa 18:7–32. Guillén, G., H. Aponte, X. Bacigalupo, and R. Rodriguez. 2015. Protozoarios de vida libre del Área de Conservación Regional Humedales de Ventanilla (Callao – Perú) en el período septiembre 2011 - enero 2012. Científica 12(1):61–69.

Young, K. 1998. El Ecosistema. In: A. Cano and K. Young (eds.). Los Pantanos de Villa: Biología y Conservación. Universidad Nacional Mayor de San Marcos, Lima - Perú. pp 3–20.

Wetland Science & Practice October 2020 245


WETLAND RESEARCH

Andes, Bofedales, and the Communities of Huascarán National Park, Peru Rodney A Chimner1, Randall Boone2, Gillian Bowser2, Laura Bourgeau-Chavez3, Beatriz Fuentealba4, Jessica Gilbert5, Javier A. Ñaupari6, Molly H. Polk7, Sigrid Resh1, Cecilia Turin8, Kenneth R. Young7, and Melody Zarria-Samanamud2

ABSTRACT ountain wetlands are abundant in the high elevations of the tropical Andes. Wetlands occupy ~11% of the total park area and are mostly found in the large mountain valleys. Wetlands occur up to 5000 m asl, but most occur between 4,000–4,700 m asl. The highest elevation wetlands are typically dominated by cushion plants, while lower elevation wetlands are more commonly occupied by graminoids. About 60% of all wetlands are peatlands and the remainder are mineral soil wet meadows. The peatlands are up to 11 m deep and 12,000 years old, storing an average of 2,101 Mg C ha-1, which is comparable to lowland tropical peatlands. Our work in Huascarán National Park in Peru is also showing the importance of wetlands in a coupled natural-human system. These wetlands and alpine landscapes are shaped in part by legacies of past human land use, including ancient pastoralism and farming, and are also affected by millions of downstream users dependent upon wetlands and glacier-fed streams for water and energy production. Biodiversity and endemism is high among taxonomic groups such as plants, birds, fish, amphibians and insects. Currently the tropical Andes are in ecological flux due to rapid land cover changes caused by both biophysical and socioeconomic drivers. In addition, the high Andes are experiencing warming and rapid glacial retreat that is resulting in hydroecological changes and socioeconomic changes to the traditional Andean societies that feed back to changes in wetland sustainability.

M

1 College of Forest Resources and Environmental Science, Michigan Technological University, Houghton, MI, USA; Corresponding author: rchimner@ mtu.edu. 2 Department of Ecosystem Science and Sustainability, Colorado State University, Fort Collins, Colorado, USA.

3 Michigan Tech Research Institute, Michigan Technological University, Ann Arbor, Michigan, USA. 4 Instituto Nacional de Investigación en Glaciares y Ecosistemas de Montaña, Huaraz, Ancash, Peru. 5 Department of Wildlife and Fisheries Sciences, Texas A&M University, College Station, TX, USA. 6 Animal Production Department, Universidad Nacional Agraria La Molina. 7 Department of Geography and the Environment, University of Texas at Austin, Austin, Texas, USA. 8 Instituto de Montaña. Lima, Peru.

246 Wetland Science & Practice October 2020

INTRODUCTION The tropical Andes are rich in biological diversity and have been utilized for millennia by human communities (Young and Lipton 2006). When the Spanish conquistadors arrived in the 16th Century, bringing horses and cattle and a different land management scheme with them, cultural resources dramatically changed and many unique ecosystems were transformed by the introduction of new grazers and browsers and the abandonment of maintenance of water systems and much of the original land use practices shifted and adapted. Dramatic transformations are once again occurring with rapid climate change and glacial retreat in the Andes of Peru that is further transforming water and biological resources, changing biodiversity as species shift, livelihoods adapt by changing agriculture, livestock husbandry, and other economic activities, and new species are introduced (Anderson et al. 2011; IPCC 2019; Hock et al. 2019). In Peru, these problems are exacerbated by growing demands for water and other mountain natural resources associated with economic growth (Mark et al. 2017; Hock et al. 2019). The high Andes of Peru have received worldwide attention due to rapid glacial retreat and socioeconomic changes to the traditional Andean societies. These changes could cause the collapse of traditional societies altering human well-being and cultural heritage, coupled with ecological collapse in one of the most biologically diverse landscapes in the world (IPCC 2019). Such dramatic shifts may be mirrored in other ecological systems and traditional societies throughout the world especially where wild pollination is tied to traditional food crops and small-scale agricultural systems (IPBES 2018; Villagra et al. 2020). A multidisciplinary research team including U.S. and Peruvian institutions, Peruvian National Park managers, agropastoralist communities and a local non-governmental organization with funding from National Science Foundation-Coupled Natural Human Systems are examining these issues in Huascarán National Park (HNP) as a case study of how communities, biodiversity, protected area management, and rapid climate change all intersect in a wetland rich mountain environment.


Natural Protected Areas (NPAs) provide critical enviBofedales (singular bofedal) is a commonly used term ronmental services in an increasingly anthropogenic world. in Peru to refer to mountain wetlands (Maldonado Fonken High elevation NPAs protect both the alpine species and 2014). One of the unique aspects of bofedales in the Andes essential watersheds needed by local communities within is the dominance of cushion plants (Figure 1). The compact and below alpine systems. Because the tropical Andes is growth form of cushion plants is an adaptation to arctic and one of the global biodiversity hotspots (Myers et al. 2000) alpine conditions as it traps heat and minimizes wind shear and high elevation endemism among taxonomic groups (Billings and Mooney 1968). In the Andes, cushion plants such as plants, birds, fish, amphibians and insects, a series are mostly in the family Juncaceae, Asteraceae and Plantagof protected areas were created to attempt to conserve that inaceae (Cooper et al. 2010). biodiversity (Rodríguez and Young 2000). HNP (340,000 ha) was created in 1975 to conserve the ecosysFIGURE 1. High elevation cushion plant peatland in Pastoruri, HNP. (Photo by Rod tems of the Cordillera Blanca, which is in the Chimner.) central Andes of Peru with high peaks ranging from 5,000 to 6,768 masl (including Peru’s highest peak, Huascarán Sur). The United Nations Educational, Scientific and Cultural Organization declared HNP a Biosphere Reserve in 1977 and included it in the list of Natural Heritage of Humanity in 1985. As of 2016, the Cordillera Blanca contained approximately 755 glaciers (25% of all tropical glaciers) and 830 lakes of glacial origin (Mark et al. 2010; Autoridad Nacional del Agua 2016). These mountain glaciers are an important source of water, and they regulate water quantity by buffering the temporal precipitation variability that supplies water for domestic, agricultural and industrial uses during the dry season for both locally and in the arid coast of Peru (Rabatel et al. 2013). HNP is surrounded by 30 agropastoralist communities that existed before park creation. Cattle and sheep are numerous and free ranging in HNP, with limited numbers of horses and FIGURE 2. Cattle and horses graze in cushion plant wet meadows and peatlands in Ulta burros allowed to graze in peripheral park areas. Valley, HNP. (Photo by Rod Chimner.) Some communities continue to have access rights to HNP land, and their animals can use grasslands and wetlands under specific agreements and arrangements with the HNP. WETLANDS Mountains are often areas of high wetland abundance due to excess water from high rates of orographic precipitation, and these wetlands provide many benefits, including high-quality habitat, nutrient sinks and transformations, carbon and water storage, and areas for pasture (Chimner et al. 2010, Cooper et al. 2012). The Andes are no exception, with wetlands common across the entire range from the northern páramo region of Colombia, Venezuela, and Ecuador through the humid and dry puna of Peru and Bolivia all the way down to southern Argentina (Chimner et al. 2011). Wetland Science & Practice October 2020 247


Cushion plant dominated bofedales are common in HNP, and many of them are organic soil peatlands or fens (Figure 1). However, other cushion plant bofedales are actually mineral soil wet meadows (Figure 2: Chimner et al. 2019). Cushion plant peatlands can have high biodiversity for species assemblages of endemics such as anurans (Lescano et al. 2020) and wet meadows often support mixed avian flocks that include rare and new-to science species (Schulenberg et al. 2020). Because cushion plant-dominated wet meadows and peatlands can look superficially

similar (Figure 2), there can be confusion in the literature and with management as they are both called bofedales (Chimner et al. 2019). While cushion plants are common, there are many other plant species found in wetlands. The most recent floristic analysis that focused on bofedales in HNP documented relatively low alpha diversity, but potentially high beta diversity. Polk et al. (2019) sampled three valleys (Llanganuco, Quilcayhuanca, and Carhuascancha) and identified 112 vascular plant species in 29 families. The most species rich families were Poaceae, Asteraceae, and Cyperaceae. The most frequent species were Plantago FIGURE 3. Graminoid peatlands and graminoid wet meadows in Rio Negro, HNP. (Photo tubulosa Decne., Eleocharis albibracteata by Rod Chimner.) Nees & Meyen ex Kunth, Juncus ebracteatus E. Mey., Gentiana sedifolia Kunth, and Calamagrostis rigescens (J. Presl) Scribn. Rarity was common in the inventory - 37% of all species occurred only once. Mean alpha diversity was 12.6, which is what might be expected for high mountain vegetation. The survey design used by Polk et al. (2019) allowed for a valleyto-valley analysis, which showed that vegetation in the three valleys are more dissimilar than they are similar, sharing only 35% of the species on average. This finding suggests that there is high beta diversity among plant species in bofedales in HNP and more vegetation surveys should be completed to further document plant diversity. Previous research identified four wetland types based on hydrophytic plant species that co-occur in Peruvian bofedales (Weberbauer 1945; Maldonado FonkĂŠn 2014). Polk et al. FIGURE 4. High elevation clear water lake in Rio Negro, HNP. (Photo by Rod Chimner.) (2019) identified five plant assemblages or groups with one cushion plant-dominated assemblage: 1) Plantago tubulosa - Oreobolus obtusangulus, and four graminiod dominated assemblages: 2) Werneria pygmaea - Pernettya prostrata, 3) Juncus ebracteatus - Carex bonplandii, 4) Eleocharis albibracteata Calamagrostis rigescens - Lachemilla pinnata, and 5) Werneria nubigena - Oritrophium limnophilum - Phlegmariurus crassus These five groups could be conceived of as subgroups of the system published previously by Weberbauer (1945) and Maldonado Fonken (2014). Cation exchange capacity, organic matter, bulk density, and elevation were factors associated with structuring bofedal vegetation in the three valleys sampled. At the scale of HNP, subsequent mapping work by Chimner et al. (2019) 248 Wetland Science & Practice October 2020


showed that elevation and latitude are additional FIGURE 5. Interdisciplinary team in the field with Peruvian and US students and local organizing factors and they classified bofedales partners in Ulta, HNP. (Photo by Molly Polk.) as cushion plant peatlands, cushion plant wet meadows (Figure 2), graminoid peatlands (Figure 3), and graminoid wet meadows (Figure 3). There is normally little information about mountain wetland abundance, because they are often small and located in remote and rugged settings that make them difficult to map (Bourgeau‐ Chavez et al. 2017). Mapping work by Chimner et al. (2019) in HNP has found that wetlands are numerous and occupy ~11% of the total HNP area, mostly found in the large mountain valleys. For context, grasslands comprise 17% and woodlands 8% of the HNP area. Cushion plant peatlands were the most abundant wetland type occupying 6.2% of HNP, followed by graminoid wet meadows (3.5%) and cushion wet meadows (1.3%), and graminoid peatlands (0.1%). Wetland type and abundance varied with elevation. Wetlands occurred up to 5,000 masl, but were most abundant between 4,000–4,700 masl FIGURE 6. Research team collaborating with community members at The Mountain with about three-quarters of all mapped wetInstitute in Huaraz, Peru. (Photo by Molly Polk.) land area occurring within this elevation zone. At the lower elevations, wetlands were mostly graminoid wet meadows and at higher elevations wetlands were mostly cushion plant peatlands. Cushion plant wet meadows were most common in mid-elevations. Wetland type and abundance also varied north to south, with more wetlands found in the south (mostly cushion plant peatlands) compared to the northern part of the park (cushion wet meadows) (Chimner et al. 2019). Wet meadows and peatlands have large differences in soils, hydrology, chemistry, and often respond very differently to grazing. Wet meadows are often preferred for grazing because they are seasonally wet and have stable soils compared with peatlands, which makes it easy for livestock to utilize. Wet meadows can still be overgrazed leading to vegetation loss, erosion, nutrient losses and soil carbon declines (Enriquez et al. 2014, 2015). In contrast to wet meadows, mountain of native species with abundant litter, high soil infiltration peatlands are very susceptible to grazing impacts due to and water quality, high plant diversity, no bare soil, and no their thick and soft organic soils, which are easily trampled sign of erosion. Whereas bofedales in poor condition had < (Chimner et al. 2010, 2011; Cooper et al. 2012) and can 25% of native species, poor soil infiltration and water qualsignificantly reduce or reverse carbon storage and increase ity, high percent cover of bare soil, and signs of erosion. emissions of the potent greenhouse gas methane by an Peatlands in HNP range in age from 3,200 to 12,000 order of magnitude (Sánchez et al. 2017). In HNP, Calvo years (Hribljan unpublished data) and have an average (2016) assessed bofedal condition in Quilcayhuanca Valley depth of ~ 5 m, with maximum depths reaching ~11 m and found that bofedales in good condition had >70% cover (Hribljan unpublished data; Chimner unpublished data). Wetland Science & Practice October 2020 249


The thick peat deposits combined with dense peat (average bulk density = 0.26 g cm-3) result in these peatlands containing very high belowground carbon stocks (average = 2,101 Mg C ha-1) compared to most peatlands globally. For instance, high elevation peatlands in HNP have carbon stocks on an areal basis that are greater than average carbon stocks of 1,421 Mg C ha-1 from peatlands in the Peruvian Amazon Pastaza-Marañon basin, which is the second largest continuous peatland complex in the tropics (Lähteenoja et al. 2012). Although little water table data exists, peatlands in HNP appear to be similar to other peatlands in that they require continuous high water tables to slow decomposition rates (Planas-Clark et al. 2020). For example, even though there were large differences in total precipitation between the wet and dry seasons, water table levels were relatively stable throughout the year and stayed near the soil surface in reference peatlands (Planas-Clark et al. 2020). This stability of water also provides fresh water during the long dry season for local agropastoralists (Maldonado Fonkén 2014). In addition to providing water, wetlands in HNP are also being used to improve water quality from heavy metal pollution from mines and glacial melting exposing geologic formations that are weathering metals. BIODIVERSITY The high Andean landscape is considered one of the world’s hotspots of biodiversity but much of that diversity is largely unknown and includes taxa that provide critical ecosystem services. High elevation wetlands and clear water lakes (Figure 4) have high diversity levels of aquatic macroinvertebrates (Nieto et al. 2020), endemic avian species (Schulenbuerg et al. 2020), amphibians (Lescano et al. 2020), and wild pollinators (IPBES 2018). Endemic species are still poorly known with new species of birds (Schulenberg et al. 2020) and insects still being discovered (Figure 5). On the landscape level, HNP provide habitat to over 780 vertebrate species, many of which are considered threatened or near threatened by the IUCN Red List of Threatened Species, including the larger animals like the puma (Puma concolor), Andean fox (Lycalopex culpaeus), and Andean bear (Tremarctos ornatus) (Fjeldså 1993; Yensen and Tarifa 2002; Lloyd 2008; Gareca et al. 2010). Many of these species are impacted by current park management. For example, the South American deer or taruka (Hippocamelus antisensis), a high altitude specialist listed as vulnerable by the IUCN Red List of Threatened Species (Barrio et al. 2017), shares habitat with domestic livestock leading to the displacement of taruka (Merkt 1987; Barrio 1999, 2006; Gazzolo and Barrio 2016). 250 Wetland Science & Practice October 2020

Biodiversity research has been limited to a few localized studies (Fjeldså 2002; Yensen and Tarifa 2002; Lloyd and Marsden 2008; Renison et al. 2018; IPCC 2019). The lack of biodiversity assessments has left gaps in understanding how change and disturbance will affect the biological communities that inhabit these systems in the future (Fjeldså 2002; Lloyd and Marsden 2008; Dangles et al. 2020). For instance, macroinvertebrates may have reduced functional diversity as glaciers retreat in high elevation systems. Most wild pollinators are poorly studied with little known about high elevation bombus species, which are of particular concern due to the introduction of non-native bombus species such as Bombus terralis. The impacts of non-native bombus introductions may be linked to the declines of the IUCN listed species such as B. dahlbomii, one of the largest bombus species (Singer and Sanguinetti 2014), and the Andean Bumble bee (B. coccineus). In addition, several species of honey producing bees such as Apis mellifera have been introduced in many high elevation communities, leading to the introduction of parasites and pathogens that appear to impact native pollinators especially, the rarer bombus species such as B. dahlbomii (deLanda et al 2020). WETLANDS AS COUPLED NATURAL AND HUMAN SYSTEMS Wetlands of the high Andes Mountains can be conceptualized as coupled natural-human systems because their characteristics and dynamics are only in part regulated by biophysical processes pertaining to the water cycle and to the dynamics of ecological succession. An important additional component is the conspicuous human dimension, such that wetland ecosystems cannot be understood without reference to interactions among the various biophysical and social processes. For instance, many Andean bofedales have been managed since pre-hispanic times by using water management technology (Morlon et al. 1996; Lane 2006, 2014). In the case of our study areas this is true even for the most remote and “pristine” parts of HNP. The human influences are ancient: one of our study sites is located along a hiking trail used by trekkers going to the Chavin archaeological site for tourism - they move along routes where pilgrims came to participate in rituals 3,500 years ago. Agropastoralists have also created and modified wetlands for livestock by diverting surface water to maintain or enlarge wetlands (Lane 2006). Similarly, farmers use canals to extend the irrigation of their crops. The sizes of these water reservoirs are 10 - 15 m in diameter and hold between 100 300 m3 of water (Lane 2014). The expansion of the wetland area not only improves the water provision service, but also regulates water flow during the dry season, increases water


quality, and contributes to soil carbon sequestration (Lane managers, and as negotiated based on past land tenure. Our 2014). Understanding the deep timeline of human influresearch suggests that the park-people interactions involve ences will be an important complement for understanding two archetypes of land use: a seasonal rotational livestock the complex couplings in this system. system that brings in cattle to the wetlands and sheep to the Inside the park today, it is not uncommon for cattle to hillslopes of the park during the dry season, and a system navigate steep slopes above 4,000 m elevation. Livestock that keeps cattle in the park year-round. The former appears are far more visible in HNP than large wild mammal speto be less environmentally damaging (and more productive cies, which are more commonly seen in landscapes outside economically), but the latter is still common because cattle HNP boundaries. The study of the pastoral systems not only owners who remove their livestock risk losing access rights provides needed information for park managers hoping in the future. The wetlands thus are viewed as critical dry to lessen damaging environmental impacts, but provides season production areas in some valleys, but can be sources information to local communities who rely on wetlands for of mortality for cattle that free range year-round, with only their economic and subsistence needs (Figures 6 and 7). weekly or biweekly visits by the livestock owners. Our This human-nature interaction exemplifies the historical research has included detailed informant interviews by cultural exchange between Western and Andean cultures, as an anthropologist fluent in Quechua, the first language of they borrowed cattle from Western culture and incorporated many local people, to get their perspectives on the costs and transformed their Andean human landscape. Cattle are and benefits of wetlands for their livelihoods. so integrated into the local cultures now that local commuIt is impossible to do research in this part of Peru and nities cannot understand life in the mountains without these not be struck by social injustices having to do with past animals. For local people, cattle are part of their natural environmental and land tenure concerns. The Santa River, landscape. It is like this in many parts of the Andes, howfed by glaciers within the national park, is littered by ever, there are still some remaining places where original unmitigated mine tailings and contaminated with heavy livestock (camelids) are the main element of the landscape. metals. More than half of the high elevations of this part Areas within HNP has been heavily used for tours by of Peru are within exploratory mining claims and several school groups; the view of glaciers and other natural fealarge mines are in active exploitation. The park provides tures was often the most direct experience Peruvian school an important protected space for natural environments, but children had of nature as seen in real life rather than on television. The glaciers proFIGURE 7. Project leaders in the field with community partners in Rio Negro, HNP. (Photo by Molly Polk.) vide critical runoff, especially during the five month long dry season, for maintaining surface water flows and for drinking water for the city of Huaraz. So, the park provides numerous educational and ecosystem services locally, regionally, and nationally. The park’s wetlands provide unique biodiversity habitat, function in ways that store water and carbon, and are conspicuous landscape features visible to park visitors. There are also important social roles filled by HNP, occupying the headwaters of valleys used by farmers for their crops downslope. The high elevations within HNP are accessible to traditional land use by the park’s neighbors through agreements made with park Wetland Science & Practice October 2020 251


must also accommodate legitimate concerns about land tenure that date back to pre-park years when lands were controlled by large landowners in the form of haciendas. Some of the concerns mentioned by local people predate the park, but still drive emotions and actions today. HNP is emblematic of global change impacts. Its boundaries include the longest extension of glaciers anywhere in the tropics. Glacier retreat is about 30% on average but on the ground, there are some places where the valley-head ice is already gone, meaning that the valley downslope is now “postglacial,” with hydrology now regulated by precipitation and groundwater, and no longer supplemented by glacial runoff. In valleys under the highest peaks, the ice is predicted to last for at least another century. So, glacial retreat is adding additional heterogeneity as to how wetlands function, and implications for human land use also vary from valley to valley depending on the size and height of glaciers’ upslope. We expect to find that this heterogeneity will be an important part of explanations for how and why wetlands change from place to place. The use of a coupled agent-based model (DECUMA) and ecosystem model (L-Range model) is being used to help us understand the complexities of these coupled natural-human interactions (Boone and Galvin 2014; Boone et al. 2018). We are simulating how landscape features such as topography, wetland location and cover, primary production, and the location of water sources influence land-use decisions of agropastoral families. The coupled model is also being used to analyze scenarios of climate change and their impacts on the landscape, wetland sustainability, and households and communities’ actions. Management scenarios such as changes in livestock population and rotation of grazing lands will be established using a participatory approach. The use of a coupled-modeling simulation approach will provide us information about the role of natural and human drivers on the ecosystem, decision-making processes, and explore the impact of short vs. long-term management options. Since wetlands are key ecosystems in this landscape, the model will provide insights of their role for land and livestock management, as well as information of how these ecosystems could be impacted due to climate change and management decisions. FUTURE ISSUES Mountain wetlands act as important sentinels of global changes, mediated by shifts in climate, associated ecological/hydrological alterations, and a suite of interactions with people. The global pandemic highlighted the intersections between large-scale landscape changes due to climate conditions coupled with changes in socioeconomic stability for local communities (loss of tourism due to COVID for 252 Wetland Science & Practice October 2020

example). Mountain wetlands are most directly affected by their hydroperiod, set by surface and groundwater hydrology, overlying biodiversity and ultimately by rainfall and runoff from glaciers in glaciated watersheds. Because of their importance for ecosystem services, they are often considered valuable by local people, who often utilize wetlands in their pastoralism systems, or by conservationists as keystone ecosystems with high endemism in protected areas such as national parks. Yet most of the ecosystem characteristics unique to the high elevation remain unknown, such as rates of endemism in high elevation birds, shifts in anuran composition in wetlands and the decline of wild pollinators such as Andean bombus species. The number of endemic species such as wild pollinators or birds is threatened by shifts in the wetlands systems coupled with the introduction of invasive pollinator species as a secondary stressor. The lack of knowledge of species diversity and overall endemism is profound especially as high elevation habitat patches shift along with local livelihoods - a perfect storm for the collapse of many ecosystem services appears to be brewing yet unseen in research inquiry. The monitoring of glacial retreat, stream discharge, and depth to water table would all provide insights into controls on hydroperiod. The monitoring of the extent and spatial configuration of the wet meadows and peatlands can be done through mapping with remote sensing (Chimner at al. 2019). In particular, our additional use of active remote sensing data through satellite-borne radar can map both surface and subsurface features. Vegetation shifts of interest in those wetlands include the shift from dominance by graminoids at lower elevations to cushion plants nearer the peaks. The human dimensions involved are multifaceted given that individual valleys in the park have different histories on land uses that are permitted, and especially in regards to the degree of control that park neighbors have (or do not have) in access rights that allow them to use park lands for livestock grazing and firewood collection. Some valleys have virtually no human impact, except for the passing of occasional hikers, while others are overgrazed with extensive trampling, an especially damaging impact on wetlands. The monitoring and evaluation of these anthropogenic influences is hence more complicated and may require nuances in mediating park-people interactions that may in some cases be antagonistic or may have histories of such antagonism. Although COVID-19 was not a consideration when we developed our research plans for the park, now it is and will require us to think about wetlands, socioeconomic drivers such as tourism, livestock and biodiversity in an entirely new context. For instance, COVID-19 has sharply reduced tourism and much of the regional interchanges of


highland agricultural products for consumer goods from the coastal cities of Peru. At HNP, the local communities have problems with their incomes since they do not have opportunities to provide guiding services, transport, hostels, and food, as well as to sell souvenirs. New approaches to food security, such as the localized production of honey, which has increased 100-fold in recent years, may also increase due to the pandemic, modifying the current land management. The local economy has been decreasing and presents an economic slowdown. The infrastructure for communications, drinking water, and public health is especially poor in smaller and more isolated settlements, perhaps meaning that subsistence farming and pastoralism is once again more important for the people living as neighbors to the park. We do not know what this will mean for park-people relations, and in particular for the ongoing dynamism of the mountain wetlands, but suggest that it be an important focus of monitoring for the next decade. n REFERENCES

Anderson, E.P., J. Marengo, R. Villalba, S. Halloy, B. Young, D. Cordero, F. Gast, E. Jaimes, D. Ruiz, S.K. Herzog, and R. Martinez. 2011. Consequences of climate change for ecosystems and ecosystem services in the tropical Andes. In: Sebastian K. Herzog, Rodney Martínez, Peter M. Jørgensen, and Holm Tiessen (eds.). Climate Change and Biodiversity in the Tropical Andes. Inter-American Institute for Global Change Research (IAI) and Scientific Committee on Problems of the Environment (SCOPE). 348 pp. Barrio, J, A. Nunez, L. Pacheco, H.A. Regidor, and N. Fuentes-Allende. 2017. Hippocamelus Antisensis, Taruca Assessment. IUCN Red List of Threatened Species. Barrio, J. 1999. Población y Hábitat de La Taruka En La Zona Reservada Aymara-Lupaca, Perú. Manejo y Conservación de Fauna Silvestre En América Latina, 453–60. Barrio, J. 2006. Manejo No Intencional de Dos Especies de Cérvidos Por Exclusión de Ganado En La Parte Alta Del Parque Nacional Río Abiseo, Perú. Revista Electrónica Manejo de Fauna Silvestre En Latinoamérica 1 (2): 1–10. Billings, W.D., H.A. Mooney. 1968. The ecology of arctic and alpine plants. Biol. Rev. 43: 481–529. Boone, R.B., K.A. Galvin, S.B. BurnSilver, P.K. Thornton, D.S. Ojima, and J.R. Jawson. 2011. Using coupled simulation models to link pastoral decision making and ecosystem services. Ecology and Society 16(2): 6. Boone, R.B. and K.A. Galvin. 2014. Simulation as an approach to socialecological integration, with an emphasis on agent-based modeling. In: M. Manfredo, et al. (eds.). Understanding Society and Natural Resources: Forging New Strands of Integration across the Social Sciences. Springer Dordrecgt Heidelberg, New York. pp. 179-202. Boone, R.B., R.T. Conant, J. Sircely, P.K. Thornton, and M. Herrero. 2018. Climate change impacts on selected global rangeland ecosystem services. Global Change Biology 24: 1382-1393. Bourgeau-Chavez, L., S.L. Endres, J.A. Graham, J.A. Hribljan, R.A. Chimner, E.A. Lilleskov, and M. Battaglia. 2017. Mapping peatlands in boreal and tropical ecoregions. Comprehensive Remote Sensing. Elsevier, Oxford. pp. 24–44.

Calvo 2016. Marco conceptual y metodológico para estimar el estado de salud de bofedales de alta montaña. Tesis de maestría Universidad Nacional Agraria La Molina Chimner R.A., D.J. Cooper and J.M. Lemly. 2010. Mountain fen distribution, types and restoration priorities, San Juan Mountains, Colorado, USA. Wetlands 30: 763-771. Chimner, R.A., G.L. Bonvissuto, M.V. Cremona, J.J. Gaitan, and C.R. Lopez. 2011. Ecohydrological conditions of wetlands along a precipitation gradient in Patagonia, Argentina. Ecologia Austral 21: 329-337. Chimner R.A., L.L. Bourgeau-Chavez, S. Grelik, J.A. Hribljan, A. Maria Planas Clarke, M.H. Polk, E.A. Lilleskov, and B. Fuentealba. 2019. Mapping extent and types of wetlands in the Cordillera Blanca, Peru. Wetlands 39: 1057-1067. Cooper, D.J., E.C. Wolf, C. Colson, W. Vering, A. Granda, and M. Meyer. 2010. Alpine Peatlands of the Andes, Cajamarca, Peru. Arctic, Antarctic, and Alpine Research 42: 19–33. Cooper, D.J., R.A. Chimner, and D.M. Merritt. 2012. Western Mountain Wetlands. In: D.P. Batzer and A.H. Baldwin (eds.). Wetland Habitats of North America: Ecology and Conservation Concerns. University of California Press, Berkeley, CA. pp. 313–328. Dangles, O., C. Ibarra, R. Espinosa, P. Andino, V. Crespo‐Pérez, D. Jacobsen, and S. Cauvy‐Fraunié. 2020. Functional structure and diversity of invertebrate communities in a glacierised catchment of the tropical Andes. Freshwater Biology, 65: 1348–1362. Enriquez A.S., R.A. Chimner, and M.V. Cremona. 2014. Long term grazing negatively affects nitrogen dynamics in Northern Patagonian wet meadows. Journal of Arid Environments 109: 1-5. Enriquez A.S., R.A. Chimner, P. Diehl, M.V. Cremona, and G.L. Bonvissuto. 2015. Grazing intensity and precipitation levels influence C reservoirs in Northern Patagonia wet and mesic meadows. Wetland Ecology and Management 23: 439-451. Fjeldså, J. 1993. The avifauna of the Polylepis woodlands of the Andean Highlands: the efficiency of basing conservation priorities on patterns of endemism. Bird Conservation International 3: 37–55. Fjeldså, J. 2002. Key areas for conserving the avifauna of Polylepis forests. Ecotropica 8: 125-131. Gareca, E.E., M. Hermy, J. Fjeldså, and O. Honnay. 2010. Polylepis woodland remnants as biodiversity islands in the Bolivian High Andes. Biodiversity and Conservation 19: 3327–46. Gazzolo, C. and J. Barrio. 2016. Feeding ecology of Taruca (Hippocamelus antisensis) populations during the rainy and dry seasons in Central Peru. International Journal of Zoology. Article ID 5806472. https://doi. org/10.1155/2016/5806472. Hock, R., G. Rasul, C. Adler, B. Cáceres, S. Gruber, Y. Hirabayashi, M. Jackson, A. Kääb, S. Kang, S. Kutuzov, Al. Milner, U. Molau, S. Morin, B. Orlove, and H. Steltzer, 2019. High Mountain Areas. In: H.-O. Pörtner, D.C. Roberts, V. Masson-Delmotte, P. Zhai, M. Tignor, E. Poloczanska, K. Mintenbeck, A. Alegría, M. Nicolai, A. Okem, J. Petzold, B. Rama, and N.M. Weyer (eds.). IPCC Special Report on the Ocean and Cryosphere in a Changing Climate. In press. IPBES 2018. The assessment report of the Intergovernmental SciencePolicy Platform on Biodiversity and Ecosystem Services on pollinators, pollination and food production. S.G. Potts, V. L. Imperatriz-Fonseca, and H. T. Ngo (eds). Secretariat of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services, Bonn, Germany. 552 pages.

Wetland Science & Practice October 2020 253


IPCC 2019. IPCC Special Report on the Ocean and Cryosphere in a Changing Climate. H.-O. Pörtner, D.C. Roberts, V. Masson-Delmotte, P. Zhai, M. Tignor, E. Poloczanska, K. Mintenbeck, A. Alegría, M. Nicolai, A. Okem, J. Petzold, B. Rama, N.M. Weyer (eds.). In press. Lähteenoja, O., Y. R. Reátegui, M. Räsänen, D. D. C. Torres, M. Oinonen, and S. Page. 2012. The large Amazonian peatland carbon sink in the subsiding Pastaza-Marañón foreland basin, Peru. Global Change Biology 18: 164–178. Lane, K. 2006. Engineering the Puna: The hydraulics of agro-pastoral communities in a Northcentral Peruvian valley. Ph.D., University of Cambridge, Cambridge. Lane K. 2014. Water Technology in the Andes. In: H. Selin (ed.). Encyclopaedia of the History of Science, Technology, and Medicine in Non-Western Cultures. Springer, Dordrecht. Lescano, J., N. Rios, G. Leynaud, J. Nori, J. Cordier, J. Lescano, N. Ríos, and G. Leynaud. 2020. Climate change threatens micro-endemic amphibians of an important South American high-altitude center of endemism. Amphibia-Reptilia : Publication of the Societas Europaea Herpetologica 41(2), 233–243. https://doi.org/10.1163/15685381-20191235 Maldonado Fonkén, M.S. 2014. An introduction to the bofedales of the Peruvian High Andes. Mires and Peat 15: Article 5. Mark, B., A. French, J. Bury, M. Carey, K. Young, O. Wigmore, M.H. Polk, J. McKenzie, L. Lautz, R. Crumley, M. Baraer, and P. Lagos. 2017. Glacier loss and hydro-social risks in the Peruvian Andes. Global and Planetary Change 159: 61-76. Merkt, J.R. 1987. Reproductive seasonality and grouping patterns on the North Andean deer or Taruca (Hippocamelus Antisensis) in southern Peru. Res. Symp. Natl. Zool. Park 388–401. Morlon, P., B. Cobo, B. Orlove, F. Palacios Rios, C. Felipe-Morales, C. Reynel, and W. Rozas. 1996. Capítulo 4. Infraestructuras agrícolas: ¿Vestigios del pasado o técnicas para el futuro? In: P. Morlon (ed.). Comprender la agricultura campesina en los Andes Centrales: Perú-Bolivia. Institut français d’études andines. doi:10.4000/books.ifea.2662 Myers, N., R.A. Mittermeier, C.G. Mittermeier, G.A. Da Fonseca, and J. Kent, 2000. Biodiversity hotspots for conservation priorities. Nature 403(6772): 853. Nieto, C., J. Rodriguez, A. Izquierdo, M. Reynaga, J. Rodríguez, and A. Izquierdo. 2020. Biological traits of macroinvertebrates from Puna peatbogs: patterns along spatial environmental gradients. Freshwater Science 39(1): 137–146. Planas-Clarke A.M, R.A. Chimner, J.A. Hribljan, E.A. Lilleskov, and B. Fuentealba. 2020. The effect of water table levels and short-term ditch restoration on mountain peatland carbon cycling in the Cordillera Blanca, Peru. Wetland Ecology and Management 28: 51-69.

254 Wetland Science & Practice October 2020

Polk M.H., K.R. Young, A. Cano, and B. León 2019. Vegetation of Andean wetlands (bofedales) in Huascarán National Park, Peru. Mires and Peat 24: 1–26. Rabatel, A., B. Francou, A. Soruco, J. Gomez, B. Cáceres, J. L. Ceballos, R. Basantes, M. Vuille, J.-E. Sicart, C. Huggel, M. Scheel, Y. Lejeune, Y. Arnaud, M. Collet, T. Condom, G. Consoli, V. Favier, V. Jomelli, R. Galarraga, P. Ginot, L. Maisincho, J. Mendoza, M. Ménégoz, E. Ramirez, P. Ribstein, W. Suarez, M. Villacis, and P. Wagnon. 2013. Current state of glaciers in the tropical Andes: a multi-century perspective on glacier evolution and climate change. The Cryosphere 7: 81–102. Renison, D, L. Morales, G.A.E. Cuyckens, C.S. Sevillano, and D.M. Cabrera Amaya. 2018. Ecology and conservation of Polylepis forests: what do we know and what do we ignore? Ecologia Austral 28 (1): 163–74. Rodríguez, L.O. and K.R. Young. 2000. Biological diversity of Peru: determining priority areas for conservation. Ambio 29: 329-337. Sánchez M.E., R.A. Chimner, J.A. Hribljan, E.A. Lilleskov, and E. Suárez. 2017. Carbon dioxide and methane fluxes in grazed and ungrazed mountain peatlands in the Ecuadorian Andes. Mires and Peat 19(20): 1-18; 10.19189/MaP.2017.OMB.277 Schulenberg, T., P. Hosner, K. Rosenberg, T. Davis, N. Krabbe, T. Schulenberg, P. Hosner, Rosenberg, T. Davis, G. Rosenberg, D. Lane, M. Andersen, M. Robbins, C. Cadena, T. Valqui, J. Salter, A. Spencer, F. Angulo, and J. Fjeldså. 2020. Untangling cryptic diversity in the High Andes: Revision of the Scytalopus [magellanicus] complex (Rhinocryptidae) in Peru reveals three new species. The Auk 137(2). Singer, R. and A. Sanguinetti. 2014. Invasive bees promote high reproductive success in Andean orchids. Biological Conservation 175: 10–20. Villagra, C., A. Vera, P. Henríquez-Piskulich, and C. Villagra. 2020. Native bees of high Andes of central Chile (Hymenoptera: Apoidea): biodiversity, phenology and the description of a new species of Xeromelissa Cockerell (Hymenoptera: Colletidae: Xeromelissinae). PeerJ., 8. https://doi.org/10.7717/peerj.8675 Weberbauer, A. 1945. El Mundo Vegetal De Los Andes Peruanos. Estación Experimental Agrícola de la Molina, Dirección de Agricultura, Ministerio de Agricultura, Lima. Yensen, E. and T. Tarifa. 2002. Mammals of Bolivian Polylepis woodlands: guild structure and diversity patterns in the world’s highest woodlands. Ecotropica 8: 145-162. Young, K.R. and J.K. Lipton. 2006. Adaptive governance and climate change in the tropical highlands of western South America. Climatic Change 78: 63-102.


WETLAND RESEARCH

Peatlands of the Central Andes Puna, South America Eduardo Oyague1 and David J. Cooper2

ABSTRACT he Andean puna is the highest elevation mountain and plateau region in the western hemisphere, extending from northern Peru to central Chile and Argentina, with vast areas above 4,250 m elevation. The Atacama Desert on the western side of the Andes is the world’s most arid region, and this aridity extends across the mountains. The upland vegetation is mostly continuous in the more humid north but becomes more and more patchy and barren composed only of bunch grasses in the central and southern region. Extensive upland talus, moraines, and colluvial deposits of the mountains produces perennial ground water flow systems that support thousands of peatlands and other wetlands that are regionally termed “bofedales” or “vegas”. The peatlands are dominated by several species in the family Juncaceae, most famously Distichia muscoides, that forms dense clonal cushions that characterize the alpine peatlands and landscapes. Being close to the equator, the growing season extends across the entire year, and continuous growth of the cushion plants have produced high carbon accumulation rates and peat bodies 7-10 m thick in many areas. The region supports unique plants and animals, and ecosystems, and the wetlands are threatened by overuse for livestock grazing, and climate changes that could reduce water provision may lead to some bofedales completely drying.

T

the world’s most biodiverse region, and on the west by the Atacama Desert, the world’s driest region. The region has complex geodynamic and volcanic processes geologic origin (Strecker et al. 2009) Above 3500 meters asl, these geodynamic factors have combined with glacial processes to shape the landscape. The central Andes alpine region is termed “Puna”, a treeless windswept area of the higher Andes, a word derived from the Quechua language. Its climate is cold and FIGURE 1. Location of Central Andes Puna ecoregions in western South America (Olson et al. 2001). A and B: Current study areas in central and southern Peru. A: Nor Yauyos – Cochas Landscape Reserve, Cordillera Central. B: Vilacota – Maure Regional Conservation Area, Cordillera del Barroso.

INTRODUCTION The Andes form the world’s longest mountain system extending more than 7000 km along the entire western edge of South America, and continuing the American Cordillera from the Arctic to the Antarctic. The mountain region from central Peru through Bolivia, northern Chile and Argentina is the highest alpine region in the western hemisphere with many peaks over 6000 meters and extensive plateaus at 3200 to 5000 meters asl. The only region on Earth with similar extensive high elevation is the Tibetan Plateau. The central Andes is bordered on the east by the Amazon basin, 1 Universidad Nacional Jorge Basadre Grohmann, Tacna, Peru; Corresponding author: eoyaguep@unjbg.edu.pe 2 Department of Forest and Rangeland Stewardship, Colorado State University, Fort Collins, Colorado, USA.

Wetland Science & Practice October 2020 255


FIGURE 2. Expansive bofedale in the highlands of southern Peru, at 4,500 m elevation. Surrounding peaks all 5200-5500 m high. (Photo by D.J. Cooper.)

semi-arid, with annual precipitation ranging from 200 to 700 mm per year. It is influenced by the movement of wet air masses from the Amazon, the temperature and currents in the Pacific Ocean, and the presence of Lake Titicaca (Falvey and Garreaud 2005)precipitation over the highaltitude plateau of the South American Altiplano exhibits a marked intraseasonal variability which has been associated with alternating moist and dry conditions observed at surface stations near the Altiplano western cordillera. In this study the characteristics of humid (wet. It is well known for having “summer every day and winter every night”, with freezing or near freezing temperatures in all seasons (Rundel et al. 1994). The puna is climatically variable, with “wet puna” in central Peru, “puna” in southern Peru and northern Argentina, and “dry puna” on the Peruvian-Bolivian Plateau called “altiplano” and in northern Chile (Figure 1). The vegetation is composed of open to continuous bunch grasses and shrubs on upland slopes with wetlands in the valley and basin bottoms. Stands of Polylepis spp. (Rosaceae) forest can occur to over 4800 m elevation on western slopes of the Andes forming one of the world’s highest

elevation forest types, for example on Nevado Sajama, the highest mountain in Bolivia. On the eastern edge of the Andes cloud forest forms the upper treeline. Many wetlands in the valleys and basins are supported by groundwater discharge and are peat accumulating fens (Cooper et al. 2010, 2019). Those peatlands are termed ‘bofedales’ in Peru, Bolivia and northern Chile, and ‘vegas’ in central Chile and Argentina (Rojo et al. 2019) (Figure 2). They have some of the most rapid known rates of peat accumulation, up to 1.4 - 2.2 mm/yr and a long term carbon accumulation rate of 37 - 47 g C/m2/yr (Hribljan et al. 2015). Carbon storage is a key ecosystem service provided by bofedales, but they also represent important refugia for many organisms, particularly during the prolonged dry seasons. Bofedales are critical pastures for indigenous pastoral communities (Yager et al. 2019) whose inhabitants live at extreme high elevations. Aymaran highlanders, who live in the puna of Bolivia, are often cited as examples of human adaptation to extreme high elevations, with blood hemoglobin oxygen levels, and lung capacity, higher than any other people. Bofedales are under intensive pressure due to direct

FIGURE 3. Intact peatland during the dry season with water table near the surface and cushions and pools separated in most areas; Apolobamba, Bolivia, 4350 m elevation. (Photo by D.J. Cooper.)

FIGURE 4. Heavy stocking rates of alpaca in bofedale in southern Peru, Vilcanota Mountains (4570 m elevation). (Photo by D.J. Cooper.)

256 Wetland Science & Practice October 2020


uses including grazing by the domesticated Andean livestock llama and alpaca, and also by European domesticated livestock particularly sheep and cattle. Many bofedales have been ditched and drained or irrigated to “improve” the pastures.

dewater bofedales. Cushion plants are sensitive to grazing. Llama and alpaca use their chisel like teeth to scrape the hard Distichia shoots from the cushions, and with heavy use the cushions are killed and replaced by short-lived and shallow-rooted plants (Figure 5). Other factors also stress these wetlands. Some industries, such as metal mining, can have significant environmental effects due to ground water extraction and pollution. The climate is also changing, as is well known from the rapid loss of glaciers throughout the tropical Andes (Rabatel et al. 2013). Some bofedales in small watersheds have completely desiccated for unknown reasons (Figure 6), while others persist but have dry season water level declines of nearly one meter (Figure 7). The hydrological functioning of Andean peatlands has been discussed and many conflicting viewpoints provided. Some have proposed their importance as reservoirs and water providers to downstream communities. Others have suggested they are dependent on glacier meltwater. However, glaciers cover < 5% of the Peruvian highlands and most fens in the puna occur in regions lacking glaciers. In

PEATLAND TYPES AND CHARACTERISTICS ‘Bofedales’ and ‘vegas’ are terms used to denotate many types of Andean wetlands, including fens, wet meadows with mineral soils, and stream floodplains (Chimner et al. 2019). However, fens are the most conspicuous and characteristic wetland in the puna. From central Peru to northern Chile fens are dominated by cushion-plants, particularly Distichia muscoides and Oxychloe andina in the family Juncaceae. In central Chile and northern Argentina fens are dominated by cushion plants with a higher presence of bunch grasses, such as Deyeuxia spp. In southern Bolivia and northern Chile and parts of Peru highly saline closed basin lakes and flats are common, including the famous Salar de Uyuni in Bolivia the world’s largest salt flat and important habitat for several species of flamingos that also inhabit many bofedales. Natural bofedales have a distinctive cushion and pool FIGURE 6. Dessicated bofedale with bare exposed peat near Huaytire, Peru structure (Figure 3). The cushion plant shoots grow so (4570 m elevation). (Photo by D.J. Cooper.) densely packed that they form an almost waterproof dam that limits lateral water movement, and when the cushions touch they isolate pools which prevents rapid drainage. Overgrazing is common as the uplands provide little forage. In addition, because there is no long lasting winter snow grazing occurs every day of the year. Stocking rates can exceed 500 animals per pasture in some districts (Figure 4). Heavy grazing can break the cushion margins allowing the pools to connect creating drainage paths that can FIGURE 5. Highly degraded cushion plant peatland with most cushions destroyed and bare ground exposed. (Photo by D.J. Cooper.)

Wetland Science & Practice October 2020 257


the puna, as elsewhere in the world, fens are groundwater dependent ecosystems, with discharge from hillslope and other aquifers and their evapotranspiration consumes water resources. The remarkable aridity of the puna ecoregion, with only 200 - 500 mm of total annual precipitation and strong seasonality with a 7 to 9 month- long dry season, reinforces bofedale dependence on perennial groundwater FIGURE 7. Dry season water table decline in peatlands dominated by Distichia muscoides cushions, Nor Yauyos – Cochas, Peru (4630 m elevation). Surrounding peaks are up to 5750 m in elevation. (Photo by D.J. Cooper.)

FIGURE 8. Interpolated groundwater isolines and dominant flow patterns in Paucarani an Oxychloe andina dominated peatland. Vilacota – Maure Regional Conservation Area, Tacna Region, southern Peru. Ground water is flowing (white arrows) from ancient glacial moraines that function as aquifer storage.

258 Wetland Science & Practice October 2020

discharge, the only water resources available during long dry periods (Falvey and Garreaud 2005)precipitation over the high-altitude plateau of the South American Altiplano exhibits a marked intraseasonal variability which has been associated with alternating moist and dry conditions observed at surface stations near the Altiplano western cordillera. In this study the characteristics of humid (wet. The stability of water provision is a key element for the ecological functionality of these ecosystems. The hydrological stability combined with the relatively constant year-around thermal regime and daylight intensity due to proximity to the equator, allows peat to accumulate (Cooper et al. 2014; Hribljan et al. 2015). These peatlands form potentially important but usually neglected natural carbon storage systems in tropical and sub-tropical regions where extensive lowland peatlands such as the Amazonian palm swamps in the Pastaza MaraĂąon, the Congo Basin, or Indonesian swamp forests were considered more important. The importance of the alpine peatlands is accentuated when considering the modelled thermal future under climate change because carbon emissions for the tropical lowland areas are predicted to rise due to higher water demands while less significant changes are predicted for the highlands (Gallego-Sala et al. 2018). PEATLAND USES AND CONSERVATION Andean peatlands are intensively used by pastoralist communities as grazing fields for native and introduced livestock and these wetlands are the only ever-green fields. That capacity for permanent use combined with the low-income of local communities mean that they must constantly increase their flock size and expose the wetlands to over use, affecting the native vegetation cover and its ecological functioning. In some native communities in central and southern Peru, Bolivia and northern Chile, which preserve traditional knowledge practices, the wetlands are managed by artificially spreading water through irrigation with the aim of increasing the grazing areas. But in the northern range of the central Puna as in Cordillera Blanca, or in the more humid eastern slopes, peatland drainage is more common than irrigation, especially in communities where traditional camelid-based livestock are replaced by introduced sheep or cattle. The peatlands are key nesting, forage and water provision habitat for native wildlife including vicuna (Vicugna vicugna), small mammals, amphibians and numerous birds including the South American species of Rhea (R. pennata and R. americana). In central Peru, cushion-plant peatlands dominated by Distichia muscoides are the only known feeding habitats for Cinclodes palliatus, a critically endangered bird species whose total population has not


surpassed 250 adult individuals in recent decades. The Peruvian government is the only central Andean country that requires priority protection for wetland ecosystems (article 99 of the General Environmental Law - Law No 28611) and considers bofedales to be fragile ecosystems needing special protection and compensation measures if disturbed. But the threats to those peatlands are varied and continue: overgrazing, illegal peat extraction, drainage, and mining and gas activities. Current and future conditions of changing climate also can affect the functionality of these peatlands changing the amount and timing of rainfall events, modifying the infiltration-runoff patterns, and altering the evapotranspiration loses.

Cordillera, one of the highest mountain ranges in Peru with tropical glaciers, a relatively humid climate with rain that varies between 400 to 700 mm/year and a seven-month dry period from May to November. Vilacota – Maure located in Tacna province, the most southern administrative region in Peru including the Atacama Desert and the dry puna includes the completely deglaciated Cordillera del Barroso as the main mountain range. The climate in Vilacota Maure is semi-arid, with precipitation varying from 200 to 400 mm/year and a long dry season from April to December. In both areas the vast majority of bofedales are cushion plant-dominated peatlands with groundwater flowing from hillslopes to bottom valleys (Figure 8) where peat thickness can be more than 10 meters. In Nor Yauyos – Cochas the dominant species is Distichia muscoides while in Vilacota Maure D. muscoides is replaced by Oxychloe andina that forms less-compact cushions and spreads more extensively reducing the presence of pools. Both in Nor Yauyos – Cochas and Vilacota – Maure the best-preserved sites with the thickest accumulated peat layers and highest organic matter content are associated with intact cushion plant communities (Figure 9). Higher diversity of plant species occurs usually in places influenced by high grazing pressure, reduced water provision, a combination of these two pressures, or in artificially created wetlands. Usually, areas with higher plant diversity have lower organic matter content and lower values of hydraulic conductivity in upper peat layers. Meanwhile, at deeper and permanently saturated levels, the peat characteristics as OM content and hydraulic conductivity as well as porosity and structure, remains relatively homogeneous, indicating possible common vegetation in the past, that has changed in modern times due to climate changes or human interventions.

OUR RESEARCH PROGRAM Our research has focused on understanding the hydrological and ecological processes that support wetlands in the Andes. We use ground water monitoring wells, piezometers, and geochemical tracers to understand the hydrologic regimes that support different wetland types and the water sources that support them (Cooper et al. 2019). The northern Andes from Colombia to northern Peru are generally much lower elevation and wetter alpine regions (termed “paramos”) where a higher diversity of peat accumulating wetland communities exist. For example, in northern Peru we identified 20 peat accumulating communities (Cooper et al. 2010). However farther south the elevation of the puna rises to much higher levels, and typically only a couple of communities occur, all dominated by cushion plants or bunch grasses. The same is true for the Andes in Colombia, where lower elevation peatlands have higher species and community richness and the highest elevations are occupied primarily by Distichia muscoides dominated cushion plant communities (Benavides and Vitt 2014). FIGURE 9. Correlation between % cover of Juncaceae (Distichia muscoiWe have also focused on the processes of peat accumu- des and Oxychloe andina) vs. % organic matter in peatlands in the Vilacota lation to determine whether in today’s climate and under – Maure Regional Conservation Area. current land uses the fens are accumulating carbon. We have used the “moss wire” method (Cooper et al. 2015), cores with 14C aging (Hribljan et al. 2015), and the use of C gas fluxes measures to demonstrate that intact cushion communities are peat forming. We are expanding into analyses of ditched, drained, and heavily grazed peatlands, as well as restored peatlands, as has been done by Rod Chimner and his students (Planas-Clarke et al. 2020). Currently, we are developing a series of research projects to understand the hydrological behavior and carbon dynamics of peatlands in Nor Yauyos – Cochas Landscape Reserve and Vilacota – Maure Regional Conservation Area, two remarkably different areas of the Peruvian puna (Figure 1, A and B). Nor Yauyos – Cochas is located in the Central

Wetland Science & Practice October 2020 259


A key research goal for the future is to quantify the condition of peatlands, their carbon stock, and the processes that sustain and degrade them. A large proportion of bofedales are highly degraded by land uses, particularly heavy and continuous grazing, and their cushion plants have been partly or completely killed. We need to understand the level of grazing that is sustainable for the persistence of ecosystem services provided by the peatlands and for the persistence of indigenous communities that rely on grazing for their livelihood. For ecosystems that are already highly degraded we need to understand how to restore their hydrologic regimes, successfully introduce cushion plants, and rebuild plant production and carbon storage. There are many challenges ahead as the people of the Andes depend on these bofedales almost completely for their livelihoods, but have used them intensively, to the point of severe degradation and collapse in some areas. The future will require intensive management and restoration to have a sustainable peatland resource in the puna. n

Falvey, M. and R.D. Garreaud. 2005. Moisture variability over the South American Altiplano during the South American low level jet experiment (SALLJEX) observing season. Journal of Geophysical Research Atmospheres 110: 1–12.

REFERENCES

Rojo, V., Y. Arzamendia, C. Perez, J. Baldo, and B. Vila. 2019. Spatial and temporal variation of the vegetation of the semiarid Puna in a pastoral system in the Pozuelos Biosphere Reserve. Environmental Monitoring and Assessment 191: 635.

Chimner, R., L. Bourgeau-Chavez, S. Grelik, J. Hribljan, A. PlanasClarke, M. Polk, E. Lilleskov, and B. Fuentealba. 2019. Mapping mountain peatlands and wet meadows using multi-date, multi-sensor remote sensing in the Cordillera Blanca, Peru. Wetlands 39: 1057-1067.

Rundel, P.W., A.P. Smith, and F.C. Meinzer. 1994. Tropical Alpine Environments. Cambridge University Press, Cambridge, U.K.

Benavides, J.C. and D.H. Vitt. 2014. Response curves and the environmental limits for peat-forming species in the northern Andes. Plant Ecology DOI 10.1007/s11258-014-0346-7.

Cooper, D.J., E. Wolf, C. Colson, W. Vering, A. Granda, and M. Meyer. 2010. Alpine peatlands of the Andes, Cajamarca, Peru. Arctic, Antarctic and Alpine Research 42: 19-33. Cooper, D.J., K. Kaczynski, D. Slayback, and K. Yager. 2015. Growth, production, and short-term peat accumulation in Distichia muscoides dominated peatlands, Bolivia, South America. Arctic, Antarctic and Alpine Research 47: 99-104. Cooper, D.J., J. Sueltenfuss, E. Oyague, K. Yager, D. Slayback, E.M.C. Caballero, … and B.G. Mark. 2019. Drivers of peatland water table dynamics in the central Andes, Bolivia and Peru. Hydrological Processes 33: 1913–1925.

260 Wetland Science & Practice October 2020

Gallego-Sala, A.V., D.J. Charman, S. Brewer, S.E. Page, I.C. Prentice, P. Friedlingstein, … and Y. Zhao. 2018. Latitudinal limits to the predicted increase of the peatland carbon sink with warming. Nature Climate Change 8: 907–913. Hribljan, J.A., D.J. Cooper, J. Sueltenfuss, E.C. Wolf, K.A. Heckman, E.A. Lilleskov, and R. Chimner. 2015. Carbon storage and long-term rate of accumulation in high-altitude Andean peatlands of Bolivia. Mires and Peat 15: 1–14. Olson, D.M., E. Dinerstein, E. Wikramanayake, N.D. Burgess, G.V.N. Powell, E.C. Underwood, … and K.R. Kassem. 2001. Terrestrial Ecoregions of the World: A new map of life on Earth. BioScience 51: 933–938. Planas-Clarke, A., R. Chimner, J. Hribljan, E. Lilleskov, and B. Fuentealba. 2020. The effect of water table levels and short term ditch restoration on mountain peatland carbon cycling in the Cordillera Blanca, Peru. Wetlands Ecology and Management 28: 51-69. Rabatel, A., B. Francou, A. Soruco, J. Gomez, B. Cáceres, J.L. Ceballos, and P. Wagnon. 2013. Current state of glaciers in the tropical Andes: a multi-century perspective on glacier evolution and climate change. The Cryosphere 7(1): 81–102. https://doi.org/10.5194/tc-7-81-2013.

Strecker, M.R., R. Alonso, B. Bookhagen, B. Carrapa, I. Coutand, M.P. Hain, … and E.R. Sobel. 2009. Does the topographic distribution of the central Andean Puna Plateau result from climatic or geodynamic processes? Geology 37: 643–646. Yager, K., C. Valdivia, D. Slayback, E. Jimenez, R. Meneses, A. Palabral-Aguilera, … and A. Romero. 2019. Socio-ecological dimensions of Andean pastoral landscape change: bridging traditional ecological knowledge and satellite image analysis in Sajama National Park, Bolivia. Regional Environmental Change 19: 1353–1369.


WETLAND RESEARCH

What is the Flora of the Pantanal Wetland? Arnildo Pott1 and Vali Joana Pott2

ABSTRACT he Brazilian Pantanal is a vast Neotropical wetland, in the upper Paraguay River basin. Rainfall and flooding are seasonal. The landscape is heterogeneous showing a mosaic of vegetation types. Overall, it is predominantly a savanna, with aquatic plants, riparian and dry forests, forest islets, woodlands, grasslands and many monodominant formations. The flora is composed of over 2,200 species of Angiosperms, and the species-richest families are Fabaceae (Pea Family) and Poaceae (Grass Family), each with over 300 species, followed by Asteraceae (Daisy Family) and Cyperaceae (Sedge Family), both with more than 100 species each. The species richest genera are Paspalum, Ipomoea (morning glories), Mimosa (sensitive plants), Croton, Eugenia, Ludwigia (primroses) and Arachis (wild peanuts). Very few are endemic species, as the region is geologically recent. The flora comes from surrounding domains, such as Cerrado, Chaco, Amazon, and Atlantic Forest, although most species have broad distribution. However, their arrangement and dynamics are unique in the Pantanal. Human population is quite low, except on the edges. Cattle ranching is the main economy for over 200 years. The conservation status of the Pantanal is still considered rather natural and pristine. Tourism is increasing, mainly for the abundant wildlife.

T

region is not a swamp, but seasonally flooded land that also includes slightly higher flood-free ground. The climate is typical of savanna, type Aw (tropical wet-and-dry), with annual rainfall around 800-1400 mm, most of which comes from variable summer rains (Arruda et al. 2016; Figure 2) creating an oscillating pluriannual flood cycle (Pott and Pott 2004), as shown in Figure 3. The irregular flood regime has been associated with oscillations of sea surface temperature such as ENSO (El Niño South Pacific Oscillation) (Thielen et al. 2020). The general flood pulse is monomodal (Junk et al. 2011), although the Paraguay River tributaries such as Aquidauana can show more than one yearly flood peak. Rainfall is higher in the upper watershed, wherefrom whatever surplus runs off to the plain. The asynchrony between local rainfall and delayed river flood creates a prolonged wet period for the riverine vegetation, allowing some Amazon rain forest species to grow there. FIGURE 1. Map of the Brazilian Pantanal wetland (grey) and the upper watershed (white) with the tributaries of the Paraguay River. (Source: Arruda et al. 2016.)

INTRODUCTION The Pantanal is the largest continuous freshwater wetland in South America, located in the middle of the continent, with 140,000 km2 in Brazil (Figure 1), and additional 15,000 km2 in near Bolivia and 5,000 km2 in Paraguay (Junk et al. 2011). It is situated in a vast intracontinental flat lowland, filled with Pliocene-Pleistocene sediments, forming a somewhat inland delta in the upper Paraguay River basin (Junk et al. 2011). It is a heterogeneous wetland for its complex hydrology (Junk et al. 2011), sediments and vegetation types, and for that reason it has been subdivided into 11 subregions. The term Pantanal, derived from pântano or swamp, can be misleading since most of the 1 Visiting Professor, INBIO-Instituto de Biociências, UFMS-Universidade Federal de Mato Grosso do Sul, Campo Grande, MS, Brazil. Corresponding author contact: arnildo.pott@gmail.com 2 Botanist, Herbarium CGMS, INBIO, UFMS, Campo Grande, MS, Brazil. Wetland Science & Practice October 2020 261


ceae, with the richest genera being Bacopa (waterhyssops), Cyperus (umbrella sedges), Eleocharis (spike rushes), Ludwigia (water primroses), Rhynchospora (beak rushes), and Utricularia (bladderworts). The Pantanal has a range of aquatic habitats and seasonally wet zones that favor a diversity of species. Most aquatic species are common to other vast South American wetlands, such as the Paraná River floodplain (extending from Brazil to the Esteros de Iberá in Argentina; Neiff et al. 2011), Amazonia (including Araguaia, Guaporé and Amapá in Brazil, and the Beni region in Bolivia; Haase and Beck 1989), and the Llanos and Orinoco basin in Colombia and Venezuela (Rial 2009). Wherefrom does the flora come? The plants of the Pantanal wetlands are composed of species from the surrounding phytogeographical provinces: Cerrado, Chaco, Amazonia and Atlantic Forest, in addition to a major contingent considered of wide distribution (Pott and Pott 2004). Cerrado, the Brazilian savanna, nearly surrounds the entire Pantanal. Chaco is a unique vegetation type in Brazil, located the southernmost subregion of the Pantanal - in Porto Murtinho, on the Paraguayan border, belonging to the Humid Chaco (Prado et al. 1992). Amazonian species reach the riparian forests because the overflow of the Paraguay River has a threemonth delay from the rainy season (i.e., it overflows three months after the local and upper watershed rainy season; Figure 2). That adds meteorological and edaphic wet periods with a similar effect of an Amazonian climate. The Atlantic Forest species include elements from the Seasonally Dry Forest. The wide-ranging group represents half of the species of the Pantanal plain (Pott and Ratter 2011; Pott and Silva 2015) and includes all plant habits, most of them grassland herbs. However, the plant diversity of the Pantanal is unique, despite the flora being made up of associations of species from various phytogeographical origins (Pott and Ratter 2011), tolerant of a harsh wet-anddry environment. The number of naturalized species is relatively high but mostly concentrated on less flooded ground. We can deduce that in a future scenario of a drier Pantanal, they will expand their distribution. Indeed, Leucaena leucocephala (leucaena or white lead tree, a fast-growing tree from Central America and Mexico) is already spreading on road embankments. So far, there are no massive exotic woody invaders that are common in other tropical wetlands, although some native shrubs (Mimosa spp. and Vernonanthura brasiliana) and trees (e.g., Curatella americana) increase in dry years. There are many naturalized herbaceous plants, mainly on less floodable grasslands. The few exotic

FIGURE 2. Mean monthly rainfall and Paraguay River level (Brazilian Navy fluviometric gauge at Ladário), showing the asynchrony of rains and flooding. (Source: Arruda et al. 2016.)

Research on the Pantanal in the last 30-40 years has been made mainly by local institutions such as Embrapa Pantanal, IPP (Research Institute for the Pantanal) and several universities (i.e., Federal University of Mato Grosso do Sul - UFMS, Federal University of Mato Grosso - UFMT, Catholic University Don Bosco - UCDB, University of the State of Mato Grosso – Unemat, and University for the Development of the State and the Pantanal Region – Uniderp), sometimes with collaboration of other Brazilian and international scientists. PLANT LIFE IN THE PANTANAL The flora of the Pantanal wetlands is represented by over 2,200 native species of Angiosperms, according to the updated checklist we compiled from reliably identified vouchers in herbaria, strictly considering the sedimentary floodplain. Nearly 800 species are illustrated in our two identification guides (Pott and Pott 1994, 2000). The most numerous families are Fabaceae (Pea Family) and Poaceae (Grass Family), each with over 300 species, followed by Asteraceae (Daisy Family) and Cyperaceae (Sedge Family), both with more than 100 species. Genera exceeding 20 species are Paspalum, Ipomoea, Mimosa, Croton, Eugenia, Ludwigia and Arachis. That follows the pattern of species of these genera being very abundant in Neotropical open vegetation types. Due to the recent geological age of the floodplain, very few species are endemic to the Pantanal, and five of them remarkably belong to the genus Arachis (wild peanuts). The number of aquatic macrophytes is over 300 species. The species-richest families are Cyperaceae and Poa262 Wetland Science & Practice October 2020


weeds to invade permanently wet habitats are Panicum FIGURE 3. Zonation of monodominant stands: anchored floating mats repens (torpedo grass) and the worst invader Urochloa arof water hyacinths (Eichhornia azurea + E. crassipes) and Pontederia rotundifolia, on the left, and emergent Oryza latifolia (wild rice), on the recta (tanner grass) (Pott and Pott 2004). right, western border of the Pantanal wetland, Brazil. (Photo by A. Pott, June Overall, most of the vegetation of the Pantanal is 2, 2009.) predominantly a seasonally flooded savanna (Pott and Pott 2004), with mosaics of aquatic plants (including floating mats, floating meadows and swamps) (Figure 3), riparian forest, dry forest, forest islets, woodlands, grasslands (Figure 4), and many monodominant formations. Plant species are arranged over a flooding gradient varying from lakes and permanent swamps to flood-free ancient levees, and floodable grassland in between (Figure 5). Soils vary in texture and fertility from 97% sandy to heavy clays, according to the sediment type. The largest alluvial fan in the Pantanal is the Taquari River, encompassing 50,000 km2 of sand. A striking feature of the Pantanal is the monodominant vegetation types, mainly floodable savannas or woodlands, with a single dominant tree, such as Attalea phalerata (uruFIGURE 4. Floodable grassland, at drawdown, Pantanal wetland, Brazil. curi palm or acuri), A. speciosa (babassu palm), Byrsonima (Photo by A. Pott, June 12, 2007.) cydoniifolia (canjiqueira), Copernicia alba (caranday palm or carandá) (Figure 6), Erythrina fusca (purple coral tree or abobreira), Handroanthus heptaphyllus (pink trumpet tree or piúva) (Figure 7) and its relative Tabebuia aurea (Caribbean trumpet tree or paratudo) (Figure. 8). Some of the monodominant species cause encroachment on natural grasslands, e.g., the trees Curatella americana (sandpaper tree or lixeira) and Vochysia divergens (cambará) (Figure 9), and shrubs as Combretum spp. (pombeiro). Some herbaceous species are also monodominant such as Cyperus giganteus (giant sedge or piri, similar to the Nile papyrus), Elyonurus muticus (carona), Oryza latifolia (wild rice), Thalia geniculata (fireflag or caeté) (Figure 10) and Typha domingensis (cattail or taboa). A very interesting vegetation type is the floating FIGURE 5. Typical landscape of the Pantanal wetland, with pond, floodable meadow (Figure 11). It is a floating islet that starts from a grassland and seasonal forest or woodland, over the topographic flooding floating mat of free-floating plants such as Salvinia augradient, at the end of flood in a wet year, Brazil. (Photo by A. Pott, May 9, riculata (water fern). On its top the sedge Cyperus blepha2009.) roleptos (formerly Oxycaryum cubense) germinates as an epiphyte and grows, forming a dense rhizomatous entangled meadow where debris and decaying plants accumulate. Within a few years a layer of floating organic soil (histosol) builds up. Caymans and birds nest on it. When that floating soil reaches 1 m deep, it is possible to walk on, and even shrubs and a few treelets grow on it. That sudd can clog old river beds and canals, or drift by winds or float downstream. When the lake dries out, the plants die, the organic soil decomposes or can be consumed by a wildfire, and the process restarts after the water returns. Interestingly, alkaline ponds (pH 8-10) called “salinas” also occur; they are surrounded by a flood-free ridge covered with dry forest and so isolated from other waters. Wetland Science & Practice October 2020 263


FIGURE 6. Monodominant stand of Copernicia alba (caranday palm or carandá), Paraguay River, Pantanal wetland, Brazil. (Photo by A. Pott, July 9, 2013.)

FIGURE 7. Monodominant floodable savanna of Handroanthus heptaphyllus (pink trumpet tree or piúva, or ipê-rosa), in the dry season, Miranda River floodplain, Pantanal wetland, Brazil. (Photo by A. Pott, September 2000.)

FIGURE 8. Monodominant floodable savanna of Tabebuia aurea (Caribbean trumpet tree or paratudo), at flood, Pantanal wetland, Brazil. (Photo by P.R. Souza, May 22, 2007.)

FIGURE 9. Monodominant floodable forest of Vochysia divergens (cambará), at flower, along seasonal streams, in flooded grassland, at drawdown, Pantanal wetland, Brazil. (Photo by Fábio Edir Costa, July 29, 2008.)

FIGURE 10. Monodominant stand of Thalia geniculata (fireflag or caeté), Paraguay River floodplain, Pantanal wetland, Brazil. (Photo by A. Pott, June 3, 2009.)

FIGURE 11. Floating meadow with 1m deep histosol (8 persons walking on including the photographer): Polygonum acuminatum in the foreground, the red plant is Rhynchanthera novemnervia, and riparian forest in the back, in the lake Baía Vermelha, near the Paraguay River, Pantanal wetland, Brazil. (Photo by A. Pott, June 2, 2009.)

264 Wetland Science & Practice October 2020


Scarce angiosperms grow inside, e.g., Ceratophyllum spp. (hornworts), besides Characeae (large algae); the water is a soup of algae, resulting in ponds of various colors. While there are no fish, amphibians, myriads of small crustaceans and water insects attract young caymans and migratory birds such as sandpipers. FACTORS AFFECTING VEGETATION We have analyzed soil seed banks from seasonal ponds, floodable grasslands, Tabebuia aurea (Caribbean trumpet tree) monodominant savanna, and riparian forest, and found that annual species predominate even under forest. The discrepant composition of the standing vegetation and the seed bank is a general finding. Under simulated flooding, aquatic plant seedlings emerged from the soil samples, while switching to terrestrial species after alternation to drained conditions. We called this a “flexible seed bank” – one that is ready to occupy gaps in either the dry or wet season quickly. Furthermore, the aquatic plants can hold the exotic grasses back (Bao et al. 2020). As a seasonal savanna, the Pantanal undergoes wildfires, often lit by lightning mainly on the surrounding hills, then fed by dry biomass of grasslands and accumulated histosol. The vegetation evolved under fire. Interestingly, even riparian forests regenerate after fires and exhibit a striking favorable interacting response to fire and flood (Arruda et al. 2016). The most fire-prone habitats are ungrazed deepflooded grasslands along the Paraguay and Miranda Rivers, compared with less flooded savannas where cattle stay year-round. When cattle are excluded, the grasslands either in swamps or dryland change to tall tussock grasses that are fuel for inevitable wildfires. In 1988, we excluded the cows from a 680 ha preserve to see what would happen: within one year the short creeping grasses were replaced by tall bunch grasses, and wildfire has swept through many times, despite the presence of a 10 m wide firebreak. The same has occurred in other areas, e.g., the 108,000 ha private preserve SESC Pantanal, created in 1996, frustrating the expectation that the grasslands would turn into a forest after cattle exclosure. Prescribed burning can be a management tool for conservation. Despite over 200 years of cattle ranching on native grasslands, the impact seems to be low because there is a surplus of available forage grass, amounts far above what can be consumed by the few species of native herbivorous mammals (Pott and Pott 2004). The most numerous of these herbivores are capybaras (Hydrochoerus hydrochaeris, the world’s largest rodent). Others are naturally more scattered, sometimes with vigorous populations despite being threatened elsewhere; they include the pampas deer (Ozotocerus bezoartius), a few related deer species ( i.e., brockets, Mazama spp.and marsh deer, Blastocerus

dichotomus), plus the tapir (Tapirus terrestris). This surplus of forage is quite different from the situation in African savannas that are under high pressure from herds of herbivores. The worst impact in the Pantanal has been from tree clearing of ridges for cultivated pastures. However, lately, deforestation has lost financing and lessened this activity in the region. Instead, many areas of coarse grasses are being replaced by Urochloa humidicola (koronivia grass – an African tropical forage species, widely cultivated in South America) around the woody patches on hummocks, thereby maintaining the general landscape pattern and even reducing fire incidence. Besides, capybara and deer relish this evergreen introduced grass. Buffaloes have a heavy impact, mainly in ponds, as we observed, reducing the species richness, in favor of a few benefiting from their dung, e.g., Pistia stratiotes (water lettuce). They are no longer raised in the Pantanal because they escape and become feral. HAVENS FOR WILDLIFE Like other sedimentary floodplains and alluvial fans, the Pantanal contains various types of water bodies, such as rivers, river branches, old river beds, oxbow lakes, lakes, permanent and temporary ponds, seasonal streams and flooded grasslands. All these water bodies are habitat for the abundant wildlife: fish, anaconda, bulldog bat, giant otter; birds such as anhinga, cormorant, ducks, herons, ibis, jabiru stork, kingfishers, limpkin, screamer, skimmer, spoonbill, terns, wattled jacana, wood stork, and many more. However, for the most part, there is no permanent water, only seasonal water bodies including excavated water holes or where underground water is pumped for cattle, and it is, of course, also utilized by the wild fauna. The harmonic coexistence of domestic and wild animals begins by sharing that provided water, as well as the salt. Despite being floodable, the Pantanal is also rich in terrestrial fauna, such as agouti (rodents), armadillos, coatimundi, crab-eating foxes, giant anteaters, howler monkeys, and peccaries. The big cats – jaguar and puma - are also frequent. Many species of birds, especially three species of macaws including the hyacinth macaw (elsewhere endangered), as well as parrots, parakeets, cardinals, the noisy speckled chachalaca, burrowing owl, flycatchers, hummingbirds, lapwing, red-legged roadrunner, rhea, toucans and woodpeckers. HUMAN USE AND CONSERVATION The Pantanal people do not hunt for meat, except the nonnative feral pig, as they prefer beef. For over 200 years, people have raised cattle in the region. The human population is quite low on the plain, except in a few towns on the edges. Crop agriculture is not allowed on the plain, Wetland Science & Practice October 2020 265


however it is naturally limited by unfavorable soils. A subsistence shift cultivation used to be practiced by riverside inhabitants and is now legally restricted under protective environmental laws. Nonetheless, land mismanagement in the 1980s (e.g., land clearing Cerrado woodland of sandy slopes for cattle pastures, creating deep gully erosion and damaging small headwater wetlands called veredas) in the upper watershed is interfering with water and soil conservation, causing significant river silting downstream, mainly in the Taquari River that has lost its bed and gallery forest and became a swamp. Silt and overflow is spreading over a nearly permanently flooded delta. Compared with the Pantanal plain, the upper watershed has a higher richness of aquatic plants, as well as of endemic species, mainly in vereda wetlands in the Cerrado savanna. These wetlands are wet year-round, functioning as water storage for the creeks that feed the rivers. There are also springs on karstic limestone of the Serra da Bodoquena plateau that support a different aquatic flora (Pott 1999). The official conservation areas are the National Park of Pantanal, the State Park Rio Negro, the State Park Encontro das Águas, and the Taiaman Biological Station (Ramsar site). Fishing is controlled, forbidden during the spawning period when river waters rise in the rainy season. Ecotourism is a growing economic activity in the region, for its abundance of wildlife and scenery, e.g. bird watching. There is also a potential for scientific tourism. In conclusion, as already stated (Heckman 1998; Junk et al. 2011), the overall conservation status of the Pantanal is still high, and it is considered a pristine, natural wetland. n REFERENCES

Arruda, W.D.S., J. Oldeland, A.C. Paranhos Filho, A. Pott, N.L. Cunha, I.H. Ishii, and G.A. Damasceno-Junior. 2016. Inundation and fire shape the structure of riparian forests in the Pantanal, Brazil. PLoS One 11: e0156825 Bao, F., T. Elsey-Quirk, M.A. Assis, E.B. Souza, and A. Pott. 2020. Do aquatic macrophytes limit the invasion of exotic species in Pantanal grasslands? Wetlands 40: 135-142. DOI: 10.1007/s13157-019-01168-5 Haase. R. and S. Beck. 1989. Structure and composition of savanna vegetation in Northern Bolivia: a preliminary report. Brittonia 4(1): 80-100. Heckman, C.W. 1998. The Pantanal of Poconé: Biota and Ecology in the Northern Section of the World’s Largest Wetland. Kluwer Academic Publishers, Dordrecht.

266 Wetland Science & Practice October 2020

Junk, W.J., C. Nunes da Cunha, C.J. da Silva, and K.M. Wantzen. 2011. The Pantanal: a large South American wetland and its position in limnological theory. In: W.J. Jung, C.J. Silva, C.N. Cunha, and K.M. Wantzen (eds.). The Pantanal: Ecology, Biodiversity and Sustainable Management of a Large Neotropical Seasonal Wetland. Pensoft Publishers, Sofia, Bulgaria. pp. 23-44. Neiff, J.J., S.L. Casco, A. Cózar, A.S.G. Poi de Neiff, and B. Ubeda. 2011. Vegetation diversity in a large Neotropical wetland during two different climate scenarios. Biological Conservation 20 (9): 2005-2020. DOI:10.1007/s10531-011-0071-7 Pott, V.J. 1999. Riqueza verde em meio azul. In: E. Scremin-Dias, V.J. Pott, R. Hora, and P.R. Souza (eds.). Nos jardins submersos da Bodoquena: guia para identificação das plantas aquáticas de Bonito e região (In the underwater gardens of Bodoquena: identification guide for the aquatic plants of Bonito and region). UFMS, Campo Grande. pp. 59-93. https://ecoa.org.org/wp - content/uploads/2016/12/Nos - Jardins - Submersos – da – Bodoquena.pdf Pott, V.J., A.C. Cervi, N.C. Bueno, and A. Pott.) 1999. Dinâmica da vegetação aquática de uma lagoa permanente da Fazenda Nhumirim, Pantanal da Nhecolândia, MS. In: Simpósio sobre Recursos Naturais e Socio-Econômicos do Pantanal; Manejo e Conservação. Corumbá, 1996. Anais. Corumbá, Embrapa. pp. 227-235. Pott, A. and V.J. Pott. 1994. Plants of Pantanal. Corumbá, Embrapa. Pott, A. and V.J. Pott. 2004. Features and conservation of the Brazilian Pantanal wetland. Wetland Ecology and Management 12: 547-522. Pott, V.J. and A. Pott. 2000. Plantas Aquáticas do Pantanal. Brasília, Embrapa. Pott, A. and J.A. Ratter. 2011. Species diversity of terrestrial plants and human impact on the vegetation of the Pantanal. In: W.J. Jung, C.J. Silva, C.N. Cunha, and K.M. Wantzen (eds.). The Pantanal: Ecology, Biodiversity and Sustainable Management of a Large Neotropical Seasonal Wetland. Pensoft Publishers, Sofia, Bulgaria. pp. 281-300. Pott, A. and J.S.V. Silva. 2015. Terrestrial and aquatic vegetation diversity of the Pantanal wetland. In: I. Bergier and M.L. Assine (eds.). Dynamics of the Pantanal Wetland in South America. The Handbook of Environmental Chemistry 37. Switzerland, Springer. DOI 10.1007/609_2015_352. Prado, D.E., P.E. Gibbs, A. Pott, and V.J. Pott. 1992. The Chaco-Pantanal transition in southern Mato Grosso, Brazil. In: P.A. Furley, J. Proctor, and J.A. Ratter (eds.). Nature and Dynamics of Forest-Savanna Boundaries. Chapman & Hall, London. pp. 451-470. Rial, A. 2009. Plantas acuáticas de los Llanos inundables del Orinoco, Venezuela. Orinoco y Amazonas editores, Caracas. Thielen, D., K-L. Schuchmann, P. Ramoni-Perazzi, M. Marquez, W. Rojas, J.I. Quintero, and M.I. Marques. 2020. Quo vadis Pantanal? Expected precipitation extremes and drought dynamics from changing sea surface temperature. PLoS ONE 15(1): e0227437. https://doi. org/10.1371/journal.pone.0227437


WETLAND RESEARCH

Connectivity of River Floodplains - The Case of Ibera Wetlands after 10,000 Years of Isolation from Parana River Juan J. Neiff1, Sylvina L. Casco1,2, Andrés Cozar Cabañas3, Alicia S.G. Poi1, Bárbara Úbeda3, Luisa F. Ricaurte4, and Eduardo M. Mendiondo5

ABSTRACT “Esteros del Ibera” is one of the most outstanding wetlands in South America, by its size (12,300 km2) and biodiversity, the largest recorded at this latitude. Owing to this, it is recognized as a Ramsar Wetland of International Importance and as a National Park in Argentina. Esteros del Ibera is an “open-air laboratory” since its landscape was created by the lateral migration of the Parana River, leaving a vast paleo-alluvial fan from Argentina to Paraguay. Here, we compare the species richness in the landscapes of Ibera with that in the equivalent landscapes of the active course of Parana River in order to understand the causes of change in diversity patterns over the past 10,000 years. We found that the loss of connectivity with the pulse regime of Parana River led to an increase in specific complexity of Ibera biota. This likely resulted from the combination of a limited change in water quality, the belonging to the vast Amazon biogeographical domain, the natural niche amplitude of wetland species, and the self-designing capacity of the Ibera ecosystem. INTRODUCTION Understanding the spatial and temporal variability of wetlands at different time scales requires analyzing the underlying biogeochemical and ecological processes (Junk 1997; Tiner 2003; Neiff 2004; Dawidek and Ferencs 2016). There is consensus on the importance of connectivity between the river’s course and the floodplain (Junk et al. 1989; Neiff 1990; Junk 1997; Melack and Forsberg 2001) as the variation in water level is responsible for a complex dynamic equilibrium in floodplain landscapes (Amoros and Bornette 2002; Thoms 2003; Bunn et al. 2006; Wiens 2009; Dawidek and Ferencs 2016). However, the effects of 1 Centro de Ecología Aplicada del Litoral (CONICET-UNNE). Corrientes, Argentina. Correspondence author contact: jj@neiff.com.ar; guadalupepoi@gmail.com 2 Facultad de Ciencias Exactas y Naturales y Agrimensura (FaCENA-UNNE). Corrientes, Argentina. sylvina.casco@exa.unne.edu.ar 3 Departamento de Biología, Instituto Universitario de Investigacion Marina (INMAR), Universidad de Cádiz, Puerto Real, 11510, Spain. andres.cozar@uca. es; barbara.ubeda@uca.es 4 Alexander von Humboldt Biological Resources Research Institute, Bogota D.C., Colombia. ricaurte.luisa@gmail.com 5 Escola de Engenharia de São Carlos Universidade de São Paulo, Brazil. e.mario. mendiondo@gmail.com

connectivity may be uneven for populations and communities (Neiff et al. 2009). Tockner et al. (1998) found that the connectivity of floodplain habitats to the river course showed diverging values for different organisms. While fish diversity was higher in the active Danube floodplain, amphibians showed greater diversity in floodplain habitats isolated from the river. Wiens (1989) pointed out that floodplain landscape should be observed from the perspective of organisms instead of from an anthropocentric viewpoint. Episodes of flooding change the proportion of aquatic and terrestrial landscapes, altering physico-chemical properties and biotic exchanges between water and land (Wiens 2002; McClain and Naiman 2008; Almeida and Melo 2009). Current concepts apply according to the particular definition of connectivity adopted. According to the River Continuum Concept (Vannote et al. 1980), the idea of longitudinal connectivity prevails. The Serial Discontinuity Concept (Ward and Stanford 1995), used for single-channel rivers, states that the stability is varyingly influenced by terrestrial ecosystems. Wiens (2002) argued that connectivity has three dimensions, namely longitudinal, lateral and vertical. The vertical dimension of connectivity seems obvious, due to the turbulence of the flow. These three dimensions of connectivity as well as the pulse regime should be evaluated at different scales depending on the landscape interactions and the processes analyzed. To the best of our knowledge, no previous study addressing the natural loss on connectivity of fluvial wetlands on the scale of thousands of years exists in the literature. We have an example from South America – for the Ibera wetland in northeastern Argentina, a former floodplain of the Parana River. • While numerous studies addressed the limnological features of Parana River and Ibera wetland since the 1980s, basic questions remain: • Was there a substitution of plant and animal species in Ibera as result of the isolation from the river’s pulses? • Has the complexity of the assemblages been modified at the level of species richness, or the spectrum of bioforms in the vegetation? Wetland Science & Practice October 2020 267


STUDY AREA Our study area includes a portion of the Parana River in northeastern Argentina, its active floodpliain and its former • Was the capture and accumulation of carbon in the floodplain - the Ibera wetland (Figure 1). Parana River is Ibera wetland modified? anastomosed in this section and runs with a slope of 0.6-0.8 We believe that studies on a long-term scale (10,000 m/km on a bed of basalts covered by poorly sorted sands and years) allow a better understanding of the resilience of clay that form bars and islands originating from the Pleistobiota, the functioning and organisation of the natural syscene to the Holocene (Orfeo and Stevaux 2002). The hydrotem. Here we provide a first attempt to address these issues logic regime of Parana River is quite irregular. It includes a based on available information. This knowledge is expected period of high waters in summer, with maximum levels in to prove useful for environmental management, biodiverFebruary-March, and a period of low waters, with minimum sity conservation, and for evaluating carbon sequestration values between August and the beginning of September. and current problems of tropical wetlands. We propose that The mean discharge is around of 18.000 m3/s and peak the collection of plant and animal species in Ibera, although flow around 65.000 m3/s. In extraordinary floods, the entire it evolved in response to the new connectivity conditions floodplain is covered by a continuous mass of water, expos(i.e., isolation from fluvial processes), preserved most of ing only the treetops of the gallery forest on the floodplain. the species of the river domain. Water of Parana River is, generally, neutral (pH FIGURE 1. Parana River and Esteros del Ibera (adapted from Poi et al. 2017). The blue 6.5-7.3), with low salinity (E.C. 40-90 µS.cm-1), line delimits the Esteros de Ibera. little calcium, abundant silica, and high turbidity and color during floods, due to loads of suspended N solids reaching 100 mg L-1 (Bonetto 1986a). PARAGUAY During the Pliocene (5.3 to 2.5 million years ago), Ibera wetland was part of the extensive active floodplain of the Parana paleo-river. River Parana River water in Ibera flowed through braided channels separated by sand bars. Movements within the Corrientes Earth’s crust during the Cretaceous Period led to sweeping changes in the regional topography, Luna Lake modifying the slope and direction of surface runoff. Paleo-river Parana gradually migrated from Galarza Lake the basin now occupied by Ibera to the current Itati Lake channel placed between Paraguay and Argentina at the end of Pleistocene roughly 10,000 years Ibera Lake ago (Castellanos 1965; Iriondo, 1991; Popolizio 1977; 2004; Neiff 2004; Orfeo and Neiff 2008; Corriente River Iwaszkiw et al. 2010). The ancient riverbed of Parana now known as the Ibera depression is now a basin of 30,000 km2 fed by rainfall (Herbst BRAZIL and Santa Cruz 1999. Currently any excess water drains into the Parana River through Corriente River in the south of the Ibera system. Its biological diversity is similar to that of the Parana River, in despite of the absence of a geARGENTINA ographical connection between them. The Iberá region was a wide fluvial belt from the Pliocene with low sinuosity braided channels separated by URUGUAY sandy natural levees. This fluvial system related with the origin of the Parana River, deposited the extense Ituzaingó Formation (fine sandy sediments with iron compounds) with outcrops in the inner part of the studied area. During the Cre• Has the isolation of the Ibera wetland produced changes in resilience?

268 Wetland Science & Practice October 2020


taceous Period, the epirogenic movements of the sub-surface basaltic blocks generated local changes of the slope, modifying the drainage direction. The Parana paleochannel gradually abandoned the central depression of Corrientes, changing to the present course that shows structural control. The fluvial belt was transformed into a series of interconnected waterbodies that receive the influence of local rains. The local climate is Humid Subtropical, with rainfall of 1500 mm/yr, athough rainfall can exceed 2500 mm in years influenced by ENSO events (Poi et al. 2017). During most of the year, temperature typically ranges from 20oC to 30oC, with a maximum absolute temperature of 44oC and minimum absolute temperature of 1oC. Rainfall has driven the geomorphology and hydrology of Ibera for millenia. No differences have been found in the pollen analyses in the whole area of study (Cuadrado and Neiff 1993), and 14-Carbon analyses gave an age of 3000 to 3700 years for the current landscape of Ibera and the older islands in Parana River. However, geomorphological and sedimentological studies agree that Ibera lost its connection with the Parana River 10,000 years ago. Therefore, the

Ibera wetland represents a unique and long-lasting environment for the collection of Amazon species that settled there (Cabrera 1951; 1976; Cabrera and Willink 1973; Carnevali 2003; Zalocar de Domitrovic 2003). It is important to note that the forms of life found in the Paleo Iberá date from 3000 to 4500 years BC (Cuadrado and Neiff, 1994; Morton, 2004; Pacella and Di Pasquo, 2020) and there are no records of plant or animal life between that date and the time of Iberá’s disconnection from the river regime. Paleontologists still do not have an answer to this absence of preserved paleontological materials, although it is known that it was a period of very contrasting climatic changes. There is a significant gap in the geological and biological history of Corrientes Province from the recent Holocene to the late Pleistocene with the 40,000-year-old Toropí Formation. Unfortunately, the sequence of the Late Pleistocene and Early Holocene has not been preserved1. Personal communication from Dr. Alfredo Zurita, Facultad de Ciencias Exactas y Naturales, Universidad Nacional del Nordeste, Argentina, aezurita74@yahoo.com.ar 1

TABLE 1. Information published about Ibera and Parana wetlands

Areas

Scientific contributions

Geology and Geomorphology

Castellanos (1965); Popolizio (1977; 1981); Herbst and Santa Cruz (1999); Iriondo (2004); Orfeo and Stevaux (2002); Orfeo and Neiff (2008)

Climatic Change effects

Neiff et al. (2011); Neiff and Neiff (2013); Úbeda et al. (2013)

Phytoplankton

Zalocar de Domitrovic (1990; 1992; 2003); Cózar et al. (2003); Zalocar de Domitrovic et al. (2007)

Zooplankton

Corrales de Jacobo and Frutos (1982); Frutos (2003; 2008; 2017); Cózar et al. (2003); Frutos et al. (2009)

Benthic invertebrates

Varela and Bechara (1981), Varela et al. (1983); Bechara and Varela (1990)

Invertebrates associated to aquatic plants

Poi de Neiff (2003); Poi de Neiff et al. (2006); Poi (2017)

Ichtyofauna

Bonetto et al. (1981); Bonetto (1986a,b); Jacobo (2002); Almirón et al. (2003), Casciotta et al. (2005); Neiff et al. (2009); Iwaszkiw et al. 2010; Contreras et al. (2017)

Vegetation

Cabrera (1976); Neiff (1986; 2003); Arbo and Tressens (2002); Carnevali (2003); Neiff and Casco (2017)

Wildlife and Biogeography

Cabrera and Willink (1973); Alvarez et al. (2003); Giraudo and Arzamendia (2003)

Ecology and Limnology

Cuadrado and Neiff (1993); Neiff et al. (1993); Canziani et al. (2003); Gantes and Torremorel (2005); Poi de Neiff (2003); Neiff (2004); Poi et al. (2017) Wetland Science & Practice October 2020 269


FIGURE 2. Satellite image of Parana River paleo-fan indicating the location of the different wetland groups identified in this study according to their connectivity to the Parana River. I: Isolated lakes and marshes; E: Eventually connected wetlands; C: Close connected wetlands.

APPROACH Forests and aquatic communities of different individual sizes and turnover rates were analyzed in the present study, ranging from phytoplankton and zooplankton to benthos and fishes. Using our own information and findings published by others over the last decades (Table 1), we compared the two wetland systems: the current floodplain of Parana River and its paleo-floodplain (Ibera) that was disconnected from the river 10,000 years ago. To detect changes at ecosystem level and their components, we use common indicators such as total number of species cited, total abundance, and dominant taxa, especially those marking functional differences. On this basis, we aim to identify which wetland “compartments” changed and which did not, how they were modified, and then try to explain “why”. We analyze whether the causes of change lie in habitat variability, the breadth of niches, or other factors. Based on our previous palynological and paleoecological research, we confirmed the fluvial origin of the current Ibera wetland system (Cuadrado and Neiff 1993; Morton 2004; Pacella and Di Pasquo 2020). In the present work, we 270 Wetland Science & Practice October 2020

try to show the biotic divergence derived from the isolation of the fluvial dynamics. We think that the indicators used here can be easily replicated in other areas as they relate to water and nature conservation management projects. Finally, we will mention what knowledge is needed based on our analysis and when it is necessary to evaluate environmental impacts of wetland isolation on a long time scale. WETLAND TYPES AND DISTRIBUTION For this region, the Parana River ecosystems contains three types of wetlands based on differences in water supply and pulse regime (Figures 2-5): isolated, eventually connected, and close connected. Isolated wetlands (Group I). They correspond to the Ibera lakes and are locally called as “esteros”. Ibera lakes are located in the fluvial paleo flatland very close to the current course of the river in the Ibera region (27°30–29°00S, 56°25–58°00W). The lakes are large (8 to 95 km2) and surrounded by extensive marshes with sudds (floating islands; Figure 3). The water was relatively transparent, slightly acidic or neutral, the conductivity ranged between 9 and 52 µS.cm-1 and the dissolved oxygen concentration varied between 5.3 and 7.5 mg/l (Neiff et al. 2011). The lakes are articulated with each other through channels of varied development, to finally resolve into a diffuse drainage system in the headwaters of the Corrientes River. This river transports water and information from the Ibera to the Parana River but it is disconnected from the hydrological pulses of this river. The water fluctuation is exclusively due to the effect of local rainfall, which is relatively predictable on annual and interannual bases. Water fluctuations are smoother (less) than in the lakes connected to Parana River. Water flows are predominantly vertical, from and to the atmosphere with strong influence from the extensive vegetation (Neiff 2004). There are laminar flows between the marshes and the large lakes and vice versa that provide a buffer against the drastic changes in the water table. The wetlands included in the Ibera depression occupy 12,000 km2. They are gently concave (0.10 to 2.5 m deep) with dense and continuous marsh vegetation covering about three quarters of total surface. General NE-SW runoff is very slow and connects with the large lakes (Galarza, Luna, Ibera, Fernández, Trin, Medina, and Itati) to finally discharge into the Parana River through the Corriente River (Figure 2). These large lakes are 2.5-4.0 m deep and the water level fluctutates from 0.5 to 0.7 m throughout the year. The western border of Ibera contains low hills of sand deposited by the Parana River and a gentle slope. Thousands of small lakes of 1-5 ha (Contreras et al. 2014) with a depth of 1.5-2.5 m are scattered across this area. Since their source of water, physicochemical characteristics, vegetation and fauna are similar to those in the large lakes of Ibera (Neiff 2004; Poi et al. 2017) they are included in Group I.


FIGURE 3. Examples of Group I (Isolated) wetlands: 1) peatland forest surrounding Galarza Lake, 2) sudds (floating islands) in Luna Lake, and 3) “esteros” landscape in Ibera Lake.

FIGURE 4. Examples of Group E (eventually connected) wetlands (from left to right): 1) marginal riparian forest, 2) bulrush marsh (Schoenoplectus californicus, Cyperus giganteus, and others) around the lakes, and 3) submerged meadows (Egeria naias, Ceratophyllum demersum and others) in lakes. Note: Inset E shows the geographic location of these wetlands. (Source of base image: Google Earth.)

Eventually connected wetlands (Group E). These wetlands, including island levee lakes, are located on ancient riverine islands. The lakes are surrounded by marshes included in high riverine islands originated by the old Parana River. These islands are near Ituzaingó city (27°31’19”S, 56°42’55”W, Figure 4). They are situated almost 3 m above the river course hence they are only connected to the river by extraordinary floods, that is, once every ten years or more. These occasional flooding events trigger an exchange of information (nutrients, organisms, seeds, eggs, etc.) between the lakes and the Parana River. Most of the time, however, the lakes are fed by rainfall. The local landscape is very similar to the Ibera region, at least in the last 3,000 years (Cuadrado and Neiff 1993). Lake waters show very low concentrations of suspended solids and a black-brown color due to the high concentration of dissolved organic matter. Close connected wetlands (C). These wetlands include shallow lakes, oxbow lakes and ponds that occur on recent lateral riverine islands that emerged in the last few centuries. They are part of the active Parana floodplain and fed by river overflows at least once a year. These waterbodies are located in the tract comprised from the wetlands in Group E to the south, at Itatí city (27°15’34”S, 58°14’ 35”W; Figure 1). Silty-sandy sediments and “white waters” (with suspended silt, fine sand and clay) predominate. The most frequent vegetation is free floating and reed swamp plants. These lakes have a high turnover of plant and animal organisms with different phases of the river pulses. The water in these wetlands is similar to that of the Parana River.

FIGURE 5. Examples of Group C wetlands: 1) Parana River floodplain with shallow and connected wetlands, 2) oxbow lakes with Pistia stratiotes floating meadows, surronded by palm forests (in the distal area of the floodplain), and 3) meander scroll covered by dense floating meadow of water hyacinth (Eichhornia crassipes).

PHYTOPLANKTON Phytoplankton composition shows relevant differences between the Ibera lakes (Group I) and the lakes of the current Parana floodplain, with ten times more species in Ibera (796 species found by Zalocar de Domitrovic 2003) than in Parana floodplain. In wetlands of Group E and C, density and diversity of phytoplankton decrease during the connection periods in relation to the disturbance and dilution produced by the river water entering into the floodplain. Wetland Science & Practice October 2020 271


Chlorophyta is the most important group in Ibera lakes (Table 2, Group I), while Bacillariophyceae was more important in the floodplain lakes of the Parana River (Table 2, Groups E and C) during the connection period (Zalocar de Domitrovic et al. 2007). Phytoplankton density in Ibera is highly dependent on the type of environment where the sample is taken, ranging from 100 to more than 4000 cells/ml. In wetlands of Parana River (Groups E and C), the density can vary between 1000-2020 (Zalocar de Domitrovic 1990; 1992) or, 5882598 cells/ml (Zalocar de Domitrovic et al. 2007) depending on the hydrological phase considered (Table 2). ZOOPLANKTON In Group I, the species richness ranged from 7 to 41 (Frutos 2003, 2017), with the highest species richness in lakes with submerged vegetation. Zooplankton abundance increased in summer (50-450 individuals/L) and decreased in winter (20-350 individuals/L). Rotifers of the genera Keratella, Ptygura, and Trichocerca were always numerically dominant (80-95%). Cladocera and copepods had variable representation in the samples (Cózar et al. 2003; Frutos 2003, 2008). The low density of cladocera and copepods in the Ibera lakes is reportedly due to high fish predation

(Cózar et al. 2003). Spatial differences in species richness were less than 38% in the Ibera lakes and the variation between high rainfall and dry periods was only 6%. The rotifers Lecane proiecta and Filinia sp. were very abundant in severe droughts. In Parana River floodplain (Group C, Table 3), zooplankton abundance is related to seasonal hydrological fluctuations, with higher concentration at the end of the lowwater seasonal period (usually in spring) and lower density during the high-water peak, due to dilution effect. Density values are often between 1 and 75 individuals/L in the river proper (Paggi and José de Paggi 1974; Corrales 1979), although these values can double in floodplain lakes (Bonetto 1986a). Rotifers are numerically dominant at all times, and cladocerans and copepods alternate as subdominants, both in considerably low abundance. Despite the difference in diversity, the dominant species are similar to those in Ibera. For instance, the most abundant rotifers include Keratella cochlearis, Trichocerca similis and Poliarthra trigla. The main difference with the lakes of Group I is the occurrence of cladocerans (e.g., Ceriodaphnia cornuta, Diaphanosoma bracyurum, and Eubosmina hagmanni) and copepods (e.g., Notodiaptomus incompositus and Mesocyclops longisetus, among others) in wetlands of Group C

FIGURE 6. Daily water fluctuation at Ibera lakes (red) and Parana River (black) measured at Itatí city from 1970 to 2019. The straight black line shows the overflow level for Ibera lakes (black), while the dashed lines show overflow levels at Parana River – the lower line indicates overflow into Group C lakes, while the upper line represent the level at which lakes in Group E are flooded with water from the Parana River. The latter lakes are rarely overflowed.

272 Wetland Science & Practice October 2020


BENTHOS The bottom substrate is an important factor influencing the benthic fauna. In Group I lakes, the bottom fauna is mainly determined by the presence of organic detritus. Since all lakes have a sandy floor, the limnetic area generally has a mobile bottom due to wind effect, so the benthic fauna is not very abundant. In the littoral zone, or in sites with dense submerged vegetation, however, the fauna is more abundant and shows a greater number of species (Varela et al. 1983; Bechara et al. 1990; Casciota et al. 2005). In Parana floodplain lakes (Groups E and C in this study), the bottom is sandy and dynamic due to the current. In the wetlands of Group C, 75 species were recorded, and in the wetlands of Group I (without connection to the river) 67 species are mentioned (Table 4). The abundance of fauna is much more variable in wetlands connected to the river course, which have a greater amount of Oligochaeta, while in Group I lakes (Ibera) Chironomidae is the dominant taxa, especially in sites with submerged vegetation. VEGETATION Of the 115 macrophytes cited for the floodplain of Parana River (Neiff 1986), only seven species are not found today in the wetlands of Ibera and they are members of the Podostemaceae - aquatic plants typically growing in the river rapids (habitats that do not exist in Ibera). These plants are only found in flowing waters and form a very specialized community known as “tachyrheophyton� (Neiff 1986). In a more recent study (Neiff et al. 2011), 161 species were catalogued for the active Parana River floodplain. They also account for 40% of the total species reported for floating islands and marshland vegetation in Ibera (400 species) based on extensive surveys and historical records in herbaria (Arbo and Tressens 2002). The comparison is especially significant since the main environmental difference between Ibera (Group I) and Parana River floodplain (Groups E and C) is related to the regime of pulses, namely amplitude, frequency and predictability of the water level fluctuations. Table 5 shows the biological spectrum of the vegetation for Groups I and C, for which information of similar amount and quality is available. The total number of species in both groups does not differ by much. The difference is the distribution of species richness in each plant bioform, which is due to the different geomorphology of both wetland groups and the variability of the pulse regime. In the Ibera depression (Group I) the habitat favors the development of emergent plants (helophytes) that have rhizomes; they dominate the marsh landscape. In Group C reed bed plants are successful because they are highly resilient to the irregular hydrological regime of the river (Neiff 1978, 1990). The number of free-floating plant species is similar

TABLE 2. Phytoplankton in Ibera lakes (grouped as I, isolated) and lakes of current Parana floodplain (grouped as E and C, eventually and closely connected respectively). Data of the taxonomic group (expressed as relative abundance (percentage of total abundance), species richness (total number of species) and abundance range (cells/ml).

Taxonomic group (%) or Other metric

I (isolation)

E and C (eventually and close connection)

Chlorophyta

65.7

49.90

Bacillariophyceae

11.93

26.95

Euglenophyta

11.80

6.50

Cyanophyta

6.91

8.51

Xanthophyceae

1.63

----

Chrysophyceae

0.75

1.63

Chryptophyta

0.75

6.51

Dinophyta Relative abundance

0.50 100%

---100%

Total abundance range (cells/ml)

140-4,033

588-2,598

Species richness

796

74

(Data Sources for Group I: Zalocar de Domitrovic 2003; for Group E: Zalocar de Domitrovic 1990 and 1992; and for Group C: Zalocar de Domitrovic 1992; Zalocar de Domitrovic et al. 2007.)

TABLE 3. Zooplankton in the lakes of Ibera (group I, isolation) and shallow lakes of current Parana floodplain (group C, closely connected). Data of the taxonomic group (expressed as a percentage of total abundance, %) and abundance (individuals/L).

Taxonomic group or Other metric

Zooplankton fauna in ecological groups according its connectivity I (Isolated)

C (Closely connected)

Rotifers

79.36

63.5

Cladocerans

14.28

17.22

Copepods

6.36

19.28

Abundance (individuals/L)

20-450

2-88

(Data Sources: CECOAL 1981; Frutos 2017.)

Wetland Science & Practice October 2020 273


in Groups I and C, and the species recorded are common to both (Table 5). However, this group of plants achieves greater coverage in the wetlands of active Parana floodplain, due to the periodic input of nutrients during the annual and more frequent floods (Carignan and Neiff 1992; Carignan et al. 1994). Although these plants can be found in Ibera, they never reach 1% coverage in the lakes. On the other hand, submerged plants form extensive meadows in the Ibera lakes, while their presence is sporadic and limited to Group E in wetlands of the Parana floodplain.

The main difference of the vegetation of Ibera in relation to that of Parana River is the area occupied by herbaceous and woody vegetation. On the islands of Parana River, forests cover about 10 to 15%, while the forest area is less than 1% in Ibera where the vegetation is virtually all herbaceous. Of the 15 tree species growing in the Parana gallery forests, only five species are found in small patches on organic (peat) or mineral soils (sand) in Ibera (Figures 7 and 8). The architecture of these trees is very different when they grow in the peaty soils of Ibera: the trees are less than 8 m high and their roots are distributed radially in the first 20 cm of the soil TABLE 4. Composition of lake bottom fauna in the study area. Data of the taxonomic group (exto avoid anoxia (Neiff and Casco 2017). pressed as a percentage of total abundance, %), species richness (total number of species) and Overall isolation of Ibera has created abundance range (individuals/m2). a different environment – a lentic one – Taxonomic group or Benthos fauna in ecological groups that supports rooted hydrophytes, while Other metric according its connectivity the Parana River favors free-floating I C plants and floodplain forests (Figures 9 (Isolated) (Closely connected) and 10). Oligochaeta 52 43 INVERTEBRATES ASSOCIATED WITH Chironomidae 41 54 AQUATIC VEGETATION Ostracoda 3 1 In extensive surveys that include several 4 2 species of aquatic and marsh plants, 152 Amphipoda+Turbelaria + morph species of invertebrates have been Acari + Mollusca recorded in Parana River floodplain (Poi 100% 100% de Neiff and Neiff 2006) and 98 in Ibera (Poi de Neiff 2003). In both surveys 5,000-10,000 1,000-100,000 Abundance (individuals/m2) identical techniques were employed on Species richness 67 75 seven of the most frequent aquatic plants (Data Sources: Varela et al. 1983; Bechara et al. 1990; Casciota et al. 2005; in wetlands (Eichhornia crassipes, EichZilli et al. 2008.) hornia azurea, Pistia stratiotes, Salvinia biloba, Azolla caroliniana, Lemna gibba, and Paspalum repens) and on five species TABLE 5. Plant bioforms in wetlands with different connectivity (expressed as species richness). in lakes of the large Ibera wetland (Typha Emerging plants are always emerging (bulrush, cattails); “Reed bed plants” have life forms adapted to flooded soil (floating rooted form) and to the emerging soil phase (emerging rooted latifolia, Leersia hexandra, E. azurea, form). Egeria najas, and Cabomba caroliniana). It is difficult to compare abundance Vegetation in ecological groups according its connectivity and composition of invertebrates to I C investigate the effects of river connectiv(Isolated) (Closely connected) ity, because different bioforms of macEmergent (cattail type) 76 38 rophytes were dominant in the Parana Reed bed plants 18 29 floodpain versus the Ibera lakes. Each plant bioform (submerged rooted, free Free-floating plants 9 10 floating, or emergent rooted) provides Rooted submerged plants 7 3 a different habitat for invertebrates. In Free submerged plants 6 3 floodplain habitat, both density and Rooted with floating leaves 6 6 species richness are influenced by horizontal flows to and from the river course. Trees on mineral or peat soils 5 15 Surveys of a floodplain lake with high TOTAL (Species richness) 127 104 connectivity to the High Parana (Sirena (Data Sources: Neiff 1986, 1990, 2003; Neiff and Casco 2017.)

274 Wetland Science & Practice October 2020


Lake with floating meadows dominated by Paspalum repens and Salvina biloba; Poi de Neiff 1981) and two Ibera lakes (Galarza and Trin with dominance of S. biloba and Eichhornia azurea; Poi de Neiff 2003) recorded 82 morph invertebrate species and 61, respectively. The comparison of similar types of habitats confirms a higher taxon richness in the lake connected to the river than the isolated lakes of Ibera. Depending on the site and the mesh size (Table 6) the overall abundance varied between 18,388 and 72,056 individuals/m2 in Ibera and High Parana, respectively. Macroinvertebrates (> 500 µm) associated with the aquatic plants were dominated by oligochaetes (mainly Naididae), insects and copepods (Table 6) both in High Parana and in Ibera. When smaller invertebrates (size greater than 125 µm) were included, copepods had the highest relative abundance in Parana and cladocerans in Ibera wetlands. Copepod species were also recorded in the plankton of the more connected lakes (Poi de Neiff 1981; Table 3). Cladocerans typically associated with vegetated areas, such as Diaphanosoma, Euryalona, Oxyurella, and Euricercus, were registered in Ibera (Poi de Neiff 2003). Mollusks and mites were poorly represented, especially in Ibera. At both sites, the composition of insects was similar. Larvae of two families Ceratopogonidae and non-biting midges (Chironomidae) were the most abundant insects. Air fronds of S. biloba supported semi-aquatic species such as the grasshopper Paulinia acuminata that has a high specificity to this host plant. There was a high number of genera of Coleoptera (Helochares, Enochrus, Derallus, Tropisternus, Paracymus, Berosus, Hydrochus, Desmopachria, Laccophylus, Liodessus, and Hydrochanthus) and Hemiptera (Belostoma, Pelocoris, Neoplea, and Ranatra) at both sites. As described above, submerged plants form extensive meadows in the Ibera lakes. The freshwater prawn Pseudopalaemon bouvieri (Decapoda) is adapted to freshwater oligohaline waters covered by submerged vegetation. It is restricted to Ibera (Group I) and other water bodies of the Corrientes province (Lopretto 1995) fed by rain; this prawn has not been reported for the Parana River floodplain (Group C). FISH FAUNA According to Bonetto (1986b), the fish fauna of this area of the Parana River contains about 200 species and does not differ much from that of other large South American rivers. As in other floodplain rivers in South America, characiforms (e.g., toothed fish) comprise almost 40% of the river fish, with many species of Tetragonoptera. Silurids (catfish) make up 20% or more of the total taxa with some being quite large fish. For example, “surubí” (Pseudoplatystoma coruscans; a long-whiskered catfish) reaches 2 m in length and may weigh 120 kg.

FIGURE 7. Gallery forest of Parana River (Group C) near of Ituzaingó: 1) high, closed, continuous forest, up to 20 m tall, and 2) Trees spaced by 4-6 m each, DBH 0.30-0.90 m; shrubs and grasses are scarce or absent as a result of frequent flooding.

FIGURE 8. Peatland forest in Group I: 1) irregular, low forest, up to 8 m high, with species as the Parana riverine forest, although with less diversity, and 2) irregularly shaped trees separated by 4-6 m (DBH 0.15-0.30m) with dense herbaceous vegetation up to 2 m high.

Some assemblages of the fish fauna are considered “sedentary fauna” - smaller fishes that live in ponds and floodplain wetlands on the islands of Groups E and C. Another group of species is the potamodromous (migratory freshwater) fishes that as adults (1-2 m long) make extensive migrations upstream in spring and downstream in late summer (Bonetto 1986). Available information shows that potamodromous species have their immature states living in lakes of Group C (Bonetto 1986b; Casciota et al. 2005; Iwaszkiw 2010; Contreras et al. 2017). These species Wetland Science & Practice October 2020 275


FIGURE 9. Parana River wetlands: 1) panoramic view of the vegetation on a meander scroll (Group C) and 2) floating meadow of water hyacinth (Eichhornia crassipes) and other species.

TABLE 6. Relative abundance of the main taxa (%) and overall abundance expressed as individuals/m2 of macro- (>500µm) and meso-invertebrates (>125µm) in the Galarza, Sirena and Trin lakes. (Sources: Reconstructed from Poi de Neiff 1981, 2003 data)

Ibera >500µm

Ibera >125µm

High Parana floodplain >500µm

High Parana floodplain >125µm

Oligochaeta

25

10.5

33

24

Cladocera

5

36

1.5

9

Copepoda

15

25

18.5

29

Amphipoda

1

0.5

4.5

26.5

Ostracoda

1

+

5

6.5

Insecta

45

18

31

18

Mollusca

+

+

1.5

0.5

Hidrachnidia

+

0.5

4

2.5

Other taxa

8

9.3

1

8

100%

100%

100%

100%

18,388

47,494

38,096

72,056

Mean overall abundance individuals/m2

The + sign indicates that the taxa were present but in a very low percentage (<0.5)

276 Wetland Science & Practice October 2020

include Prochilodus lineatus, Pterodoras granulosus, Oxydoras kneri, Trachydoras paraguayensis, Serrasalmus spilopleura, S. marginatus, Pygocentrus nattereri, Hypostomus latifrons, Loricariichthys melanocheilus, Schizodon borelli, Leporinus lacustris, Pachyurus bonariensis, Triportheus paranensis, Odontesthes perugiae, and Potamotrygon motoro. Some larger fish (Salminus brasiliensis, Pseudoplatistoma coruscans, P. fasciatum, and Lucypimelodus pati) of the Parana River are migrants each year during high flows that occasionally also go upstream via the Corriente River to the southern lakes of the Ibera (Itatí, Medina lakes); they are, however, not found in the isolated lakes of northern Ibera (Galarza, Luna, and Ibera lakes). In Group E lakes, fishes associated with littoral, emergent or rooted floating vegetation are very common (Iwaszkiw et al. 2010). These fishes include Poptella paraguayensis, Hyphessobrycon eques, Moenkausia spp., Acesrorhynchus pantaneiro, Hypostomus latifrons, Hypoptopoma inexpectata, Cichlasoma dimerus, Gymnogeophagus balzanii, Apistogramma spp., Crenicichla spp., Hoplerythrinus unitaeniatus, Hoplias malabaricus, Hoplosternum littorale, and Lepthoplosternum pectoral. Many of these fish are also found in Group I where the most frequent fishes are Hyphessobrycon eques, Moenkausia spp., Acesrorhynchus sp., Cichlasoma dimerus, Gymnogeophagus balzanii and Apistogramma spp. The migratory fish of Parana River (Salminus brasiliensis, Pseudoplatistoma coruscans, P. fasciatum; Lucypimelodus pati) are of occasional presence, as they enter some lakes only during the extraordinary floods of Parana River. As shown in Table 7, fish fauna appears similar across the region regardless of connection to the river. Although the percentage of characiforms is higher than that mentioned by Bonetto (1986b), the percentage of species included in the different taxonomic groups has a similar proportion in the three connectivity vari-


ants (Groups I, E and C). Of the 200 species reported by Bonetto (1986a) for the Upper Parana River, there are 111 species in Ibera (Casciotta et al. 2005). This total accounts for those upstream-migrating species found in lakes and streams of the southern Ibera basin (Group I) which are still linked to Parana River, at south of Itatí lake (Figure 2). However, in the lakes of northern Ibera, which are completely isolated from the pulses of the Parana River (lakes Galarza, Ibera and Luna), the same authors reported only 51 species, reflecting the impact of isolation from the river. The loss of connectivity of the wetlands of Ibera (Group I) has resulted in a reduction of the number of fish species to a third or a quarter and a loss of species of large migratory (potamodromous) fish that are restricted to the Parana River and its active floodplain wetlands.

exclusive to Group I. Perhaps time (3,000-4,000 years) was not sufficient or the differences in environmental conditions (e.g., climate and water quality), except for flooding and soil saturation, were not significantly different between the Groups. For some components of the system (Group I) such as phytoplankton, there was even an increase one order in magnitude in species richness, due to the lower rate of water renewal resulting from lower water level fluctuations. Despite this, there is no unique species in Group I. All the species recorded there were found in other environments in the Parana River basin. In the Group I, more than 80% of the marsh area is occupied by plants with rhizomes (helophytes). This type of macro-vegetation in the lacustrine scenario is surprisingly not affected by flooding because the entire marsh with its peaty soils simply floats, rising with the increase in water level (i.e., floating islands called “embalsados”) and falling when water levels decline. The marsh vegetation is sensitive to prolonged dry periods, yet since most of the species are helophytes with rhizomes that are able to root in the substrate, allowing the plants to persist during the dry period. Most of the forests species in Parana River area are adapted to flooding periods. The floodplain trees lack adaptations for survival during periods of prolonged dryness. This fact likely explains both the small area of forest and its low species richness.

COMPARISON OF FLUCTUATION REGIMES UNDER DIFFERENT CONNECTIVITY 10K YEAR LATER Daily data on water level fluctuations have been collected since 1929. The fluctuation of the water level in wetlands of Group I has maintained a pulse regime as a consequence of the seasonality of the rains (Neiff 2004). The range of fluctuation between maximum and minimum absolutes was close to 1.5 m (Figure 6). The pulse rate in Parana River is very irregular and the extreme range of fluctuation between highs and lows is greater than 8 m historically. When the water level reaches around 3.5 m, Parana River spills over into many of the lakes classified as Group C, receiving new water and information. If the riverwater level exceeds 9 m line, the lakes of Group E are then flooded; since 1970, this happened in 1983 and FIGURE 10. Lentic habitats of Ibera wetlands (Group I): 1) view of open water of the Ibera Lake again in 1992 (Figure 6). and 2) marsh vegetation in littoral zone. The pulse regime in wetlands indicates horizontal movements of water from the river and allows some horizontal circulation of information (nutrients, eggs, seeds, plankton, etc.). Water turnover produces renewal in biotic assemblages through water circulation but also through change in habitat conditions (e.g., transparency, oxygen, and nutrients). In Group I (total isolation; at least for lakes at north of Itatí Lake) there is no exchange of information with the river, nor is there a washout effect, and less variability of the habitat is maintained, favoring the permanence of a greater number of species, with the taxonomic configuration of Amazonian origin, especially in the plankton, benthos, aquatic and marsh vegetation. Interestingly, the isolation did not manage to produce any endemic species

Wetland Science & Practice October 2020 277


ecosystems (Cabrera and Willink 1973) and, currently, as an ecological corridor Fish fauna in ecological groups according Taxonomic group for many species (Bonetto 1986a; 1986b; its connectivity Giraudo and Arzamendia 2003), it also C appears to have functioned as a meridian I E (Closely barrier to the dispersal of some birds as (Isolated) (Eventually connected) connected) Thamnophilus caerulescens, Cyclarhis Characiforms 66 46 63 gujanensis, Thraupis sayaca, L. angustirostris, and Colaptes melanochloros Siluriforms 15 27 23 (Kopuchian et al. 2020). According to Perciforms 10 16 9 these authors, large rivers as Parana, Gymnotiforms 4 4 3 function as a barrier to genetic flow in a transverse direction between both banks Cyprinodontiforms 4 4 for some terrestrial birds, leading to Pleuronectiforms 2 1 population differentiation and, ultimately, Beloniforms 1 1 1 allopathic speciation. Isolation changed Total 100 100 100 the landscape pattern with greater variety of habitat (large lakes, marshes, peat(Data Sources: Casciota et al. 2005; Iwaszkiw 2010; Contreras et al. 2017.) land areas, running waters, and riparian forests), and greater species richness in DISCUSSION some communities (plankton, benthos, and aquatic plants) After 3,000-4,000 years, it appears that the natural system and simplification in others (disappearance of migratory of Ibera has maintained a part of its original configuration fish and some tree species from the riparian forests). and adapted some elements and processes to a new habiThe increased complexity that arose in Ibera from tat configuration, according to its capacity of self-design isolation can be explained from different perspectives: 1) (Mitsch and Jørgensen 2003; Odum and Odum 2003). the biogeographical context, 2) the extent of the niche of Since then Ibera has remained isolated from the Parana the resident species, 3) the nature of change in the enviRiver with changes in the biota due to dry period climate changes in the Lower and Middle Holocene, and a progres- ronment, and 4) the ability of the system for self-design. From the first point, we consider that the impact of loss of sively wetter period in the last three or four thousand years in the recent Holocene. Previous studies have demonstrated floodplain connectivity to the river becomes particularly significant if the Ibera wetland system constitutes a speciesthat the Parana River fed Esteros del Ibera produced a similar pattern of the landscape that was maintained over at endemic area. For this system, all species belong to the vast Amazon domain, which has remained a biogeographic least three thousand years ago for the entire Group I (Neiff dispersal center even during the glacial period. Since the 2004; Orfeo and Neiff 2008; Pacella and Di Pasquo 2020). Although sedimentological and geomorphological evidence Amazon bioprovince is among the most species-rich in the world, the lack of endemics is not surprising especially has shown that the Ibera marshes (Group I) were isolated from the Parana River 10,000 years ago (Castellanos 1965; when wetland species generally have very broad niches. Orfeo 2005; Orfeo and Neiff 2008; Orfeo et al. 2014), paly- Despite this, the ecology of the Ibera has changed comnological information shows that the current landscape cor- pletely as a consequence of its isolation, with a greater extension and variety of lentic environments, although responds to a recent humid tropical phase from 3000-3500 frequent species have disappeared from the river habitat of years old (Cuadrado and Neiff 1994; Pacella and Di Pasthe Parana. quo 2020). The same authors agree that there is no pollen The breadth of the ecological niches also plays a key evidence of older landscapes. In this contribution we point role in the pathway for the ecological change. Since the out that, although Esteros del Ibera was isolated from the Upper Pliocene, the climate of the subtropical zone of Parana River for 10,000 years, experiencing long periods South America has gone through very contrasting wet and of dry and wet weather, they were able to maintain a fairly dry periods (Iriondo 2004), driving a selection of plants and similar assemblage of plant and animal species although animals adapted to the irregular water regimes and drastic not all species were able to adapt to the new conditions. changes in the configuration of their habitat. Thus, an imWhile the Parana River in its north-south direction has portant number of fluvial species remain in the Ibera lakes served as a vector for the dispersal of genetic information (Group I) after 10,000 years of isolation and there is no refrom the Amazon region to Parana River and surrounding TABLE 7. Percentage of fish fauna represented by taxonomic group for the three wetland groups.

278 Wetland Science & Practice October 2020


cord of species unique to Ibera. An example of the ecologi- lands changed from a lotic system dominated by river overcal plasticity of the wetlands in the region occurred in 1995, flows to a lentic system of lakes with fluctuations due to when the Yacyretá Reservoir was constructed by covering local rainfall patterns. It is evident, for example, that Ibera the marshes of ancient islands in Parana River (Group E) wetlands (in Group I) have a greater number of Cyperaceae with an eight-meter-thick water layer. After seven months and other helophytes unlike riverine wetlands (Group C) of being completely submerged, extensive “islands” of that have a greater occurrence of free-floating plants. The several kilometers of peat rose to the surface. Only 22 days species of trees that live in Ibera are smaller than those on later, the herbaceous and shrubby vegetation had sprouted the islands of Parana River and have a highly developed from the buds of the plants that were lying on the surface root system in the shape of a big dish, which allows them to looking “dead”. It is expected that the floating islands of be supported on organic soils (peat). The process of selfIbera formed in a similar manner, thereby maintaining most design seems to have selected species that are extremely of the plants found on the former Parana floodplain. tolerant of climatic change, which means that the structure Ibera has experienced strong climate disturbances in of the Ibera landscape and its biotic components are mainthe past, and global climate change is expected to have an tained with a low rate of change in spite of extreme climatic impact on the region. We studied the possible effects on events of drought and extraordinary waterlogging (Neiff the Ibera lakes (Group I) under two future climatic sceet al. 2011; Ubeda et al. 2013). Yet, at the same time, the narios (A2 and B2) proposed by the International Panel on lower variability of the water sheet and the lower flow rate Climate Change (IPCC; Neiff and Neiff 2013; Ubeda et has favored the increase of numerous species of plankton al. 2013). Our results suggest that even though a reduction and benthos that take advantage of the microenvironments in lakes size could have negative effects on biota, affectof the waters with different types of vegetation. ing richness species at local scale (Ubeda et al. 2013), the The P/R ratio is higher in Ibera wetlands than in the biodiversity will not be significantly affected (Neiff et al. wetlands of Parana River resulting in the accumulation 2011; Neiff and Neiff 2013; Ubeda et al. 2013). The vastof organic matter in the Ibera marshes (“esteros”) that ness of Esteros of Ibera wetland complex, with its huge surround the lakes. This organic matter, although slowly variety of habitat types and ample niches for most species degraded, releases substances that are recycled by the vegmake it highly resilient from the biodiversity standpoint. etation of the lakes. We believe that knowledge of the breadth of niches and the At least in the time scale of our analysis, biodiversity does resilience of the landscape are key aspects for the scientific not seem to be a powerful indicator to evaluate the effect of the assessment of global climate change risks. The pulse regime has become more regular, FIGURE 11. Seasonal changes in the water levels for the Ibera wetlands (Group I) and the floodplain of Parana River (Group C). The meters represent the vertical variability of showing shorter range of water fluctuations the water sheet in each gauge station. In Group I the seasonal fluctuation is lower than in after isolation from the river (Figures 6 and Group C because it responds to the variation of local rainfall that occurs over a wide area. 11). Water in Ibera comes from rainwater that The vegetation of the marsh increases the roughness of the surface and decreases the has been draining through sand for thousands runoff speed. of years. However, the quality of Ibera waters is not very different from the waters of Parana River: low electrical conductivity (EC), slightly acid to neutral pH, low nutrient content (especially nitrogen) and the ionic balance is of the type: HCO3-> Na+> Cl-> Mg+> SO4-> Ca+> K+. The water exchanges between the marshes and Ibera lakes determine the contribution of chemical substances from the organic soils (Neiff 2004; Ubeda et al. 2013; Poi et al. 2017). Ibera’s self-designed response to loss of connectivity is interesting. Nature has selected those tree species that can persist in water-saturated soils for long periods and go through prolonged dry spells - trees that can occupy loose, acidic soils (pH 4.5), such as peat or sandy soils with very low nutrient concentration. Ibera wetWetland Science & Practice October 2020 279


loss of connectivity. We know that all the biota of Parana River comes from the Amazon mega-basin where extensive floodplains have remained since geological-evolutionary times. So what biotic changes have occurred? The dominance of rotifers in the potamoplankton could be a consequence of isolation. Loss of connectivity with Parana River during Upper Pleistocene produced wetlands with an increase in different forms of organic matter, which is the favorable habitat for rotifers. We think that the analysis of connectivity in terrestrial ecosystems based on topological relationships (as the spatial proximity of landscapes or as exchanges between populations or landscape gene banks) is inappropriate in river systems, due to differences in the response mechanisms of individual organisms and landscapes to changes to the pulse regime and variables associated with river connectivity. While in a terrestrial native forest the loss of landscape continuity is seen as fragmentation and increased distance between patches, in a floodplain of large rivers it is normal for the natural design of the landscape to include numerous patches of forest in the form of “galleries” or “patches” in the landscape matrix due to differences in the topographical position. This determines different eco-hydrological connectivity and, consequently, different assemblages of species, such as separate cells in the landscape. We have a difficult challenge in the study of the niches (in Hutchinson`s sense) of plants and animals, in order to assess the relationships of river connectivity in different scales of time and space. Although the hydrological regime is one of the main characteristics that condition and define the functioning of aquatic systems, there are other attributes that determine the character of wetlands, as Tiner (2017) pointed

out. In the case of Ibera, there was a drastic change in the hydrological regime with a significant attenuation of the variability at interannual and seasonal scales (Figures 6 and 11). This change resulted in increased species richness for most biotic assemblages over the millennia time scale (plankton, benthos, and aquatic vegetation), except for forests where the number of species was less than ithe Parana gallery forests.

CONCLUSION Generally, the connectivity of basin landscapes is analyzed on a current scale, or that of the recent past, without putting into a biogeographical context the events and changes that occur in the support system (physical-chemical environment), focusing the analysis on the effects of engineering works on the stability of riverine wetlands (e.g., damming of rivers or channelization of watercourses). Undoubtedly, any human action on the ecosystems produces disturbances that can alter the local, regional or global nature in different ways. Our challenge is to understand the impact of natural changes in connectivity between landscapes in a basin. We have compared a scenario in which the isolation of riverine wetlands occurred naturally 10,000 years ago. While the isolation of Ibera has clearly created a lentic environment on the former floodplain and a decline in fish species, this species segregation affected only the northern lakes of Ibera, because large migratory fish need to migrate in order to reproduce. We have not recorded the appearance of unique species typical of the new situation of isolation. Obviously, the absence of rapids in Ibera justifies the disappearance of typically rheophile species such as those of the Podostemaceae family cited for Parana River (Neiff 1986). Nor have we recorded in Ibera the presence of any invasive species that occur in the Parana River in recent decades, FIGURE 12. Synopsis of the changes produced by the disconnection of “Esteros de Ibera” such as the golden mussel (Limnoperna fortufrom the water regime of Parana River. nei) or the tilapia (Tilapia niloticus). We can think that the structure and functioning of the Ibera macro-wetland has retained its biodiversity and as a result of isolation, it will resist biological invasions, at least from riverine species. As a synthesis we present Figure 12 with the most notable changes in the long-term scale. n REFERENCES

Almeida, F.F. and S. Melo. 2009. Limnological considerations about an Amazonian floodplain lake (Catalão lake—Amazonas State, Brazil). Acta Scientiarum Biological Sciences 31: 387–395. https://doi.org/10.4025/actascibiolsci.v31i4.4641. Almirón, A.E., J.R. Casciotta, J.A. Bechara, J.P. Roux, S. Sánchez and P. Toccalino. 2003. La ictiofauna de los esteros del Iberá y su importancia en la designación de la reserva como Sitio Ramsar: In: B. Alvarez (ed.). Fauna del Iberá. Editorial de la Universidad Nacional del Nordeste, Corrientes. pp. 75-85. 280 Wetland Science & Practice October 2020


Alvarez, B.B. (ed.). 2003. Fauna del Iberá. Editorial de la Universidad Nacional del Nordeste, Corrientes.

climáticos e hidrológicos en el area del Embalse Paraná Medio Norte. CECOAL: 267-293.

Amoros, C. and G. Bornette. 2002. Connectivity and biocomplexity in waterbodies of riverine floodplains. Freshwater Biology 47: 761–776. https://doi.org/10.1046/j.1365-2427.2002.00905.x.

Cózar, A., C.M. García, and J.A. Gálvez. 2003. Analysis of plankton size spectra irregularities in two subtropical shallow lakes (Esteros del Iberá, Corrientes, Argentina). Canadian Journal Fisheries Aquatic Science 60: 411-420.

Arbo, M.M. and S.G. Tressens (eds.). 2002. Flora del Iberá. Editorial de la Universidad Nacional del Nordeste, Corrientes.

Cuadrado, G.A. and J.J. Neiff. 1993. Palynology of embalsados in distrophic lakes in Northeastern of Argentina. Revista Brasileira de Biologia 53: 443-451.

Bechara, J.A. and M.E. Varela. 1990. La fauna bentónica de lagunas y cursos de agua del sistema Iberá (Corrientes, Argentina). Ecosur 16: 45-60. Bonetto, A.A. 1986a. The Paraná River System. In: B.R. Davies and K.F. Walker. The Ecology of River Systems. Dr. W. Junk Publ., The Netherlands. pp. 541-555. Bonetto, A.A. 1986b. Fish of the Paraná system, In: B.R. Davies and K.F. Walker. The Ecology of River Systems. Dr. W. Junk Publ., The Netherlands. pp. 573-588. Bonetto, A.A., M. Canon Verón, and D. Roldán. 1981. Nuevos aportes al conocimiento de las migraciones de peces en el río Paraná. Ecosur 8: 29-40. Bunn, S.E, M.C. Thoms, S.K. Hamilton, and S.J. Capon. 2006. Flow variability in dryland rivers: boom, bust and the bits in between. River Research and Applications 22: 179–186.

Dawidek, J. and B. Ferencz. 2016. Historical changes of hydrological connectivity of selected polish floodplain lakes. River Research and Applications. 32: 1862–1871. Frutos, S.M. 2003. Zooplancton de lagunas y cursos de agua del Sistema Iberá. In: A. Poi de Neiff (ed.). Limnología del Iberá. Aspectos físicos, químicos y biológicos de las aguas. Editorial de la Universidad Nacional del Nordeste, Corrientes. pp. 143-161. Frutos, S.M. 2008. Biodiversidad del zooplancton en Corrientes, Chaco y Formosa. In: S.L. Casco (comp.), I Basterra, and J.J. Neiff (dir.) Manual de Biodiversidad de Corrientes, Chaco y Formosa. Editorial de la Universidad Nacional del Nordeste, Corrientes. pp. 79-92.

Cabrera, A.L. 1951. Territorios fitogeográficos de la República Argentina. Boletín de la Sociedad Argentina de Botánica 4: 1-65.

Frutos, S.M. 2017. La vida suspendida en el agua: la diversidad del zooplancton en los ambientes acuáticos del Iberá. In: A.S.G. Poi (comp.). Biodiversidad en las aguas del Iberá. Editorial de la Universidad Nacional del Nordeste, Corrientes. pp. 99-118.

Cabrera, A.L. 1976. Regiones fitogeográficas argentinas. Enciclopedia Argentina de Agricultura y Jardinería. ACME, Buenos Aires.

Gantes, P. and A. Torremorel. 2005. Production and decomposition in floating soils of the Iberá wetlands. Limnetica 24: 203–210

Cabrera, A.L. and A. Willink. 1973. Biogeografía de América Latina. OEA. Washington D.C. Serie de Biología, Monografía 13.

Giraudo, A.R. and V. Arzamendia. 2003. ¿Son las planicies fluviales de la cuenca del plata, corredores de biodiversidad? In: J.J. Neiff (ed.). Humedales de Iberoamérica. Programa Iberoamericano de Ciencia y Tecnología para el Desarrollo (CYTED), pp. 157-170.

Canziani, G., C. Rossi, S. Loiselle, and R. Ferrati. 2003. Los Esteros del Iberá. Informe del Proyecto El Manejo sustentable de Humedales del Mercosur Fundación Vida Silvestre, Buenos Aires. Carignan, R. and J.J. Neiff. 1992. Nutrient dynamics in the floodplain ponds of the Paraná River (Argentina) dominated by Eichhornia crassipes. Biogeochemestry 17: 85-121. Carignan, R., J.J. Neiff, and D. Planas. 1994. Limitation of water hyacinth by nitrogen in subtropical lakes of the Paraná floodplain (Argentina). Limnology and Oceanography 39: 439-443. Carnevali, R. 2003. El Iberá y su entorno fitogeográfico. Editorial de la Universidad Nacional del Nordeste. Corrientes. Carnevali, R.P., P. Collins, and A.S.G. Poi. 2016. Reproductive pattern of the freshwater prawn Pseudopalaemon bouvieri (Crustacea, Palaemonidae) from hypo-osmotic shallow lakes of Corrientes (Argentina). Studies in Neotropical Fauna and Environment 51: 159-168. Casciotta, J.R., A.A. Almirón, and J. Bechara. 2005. Peces del Iberá. Hábitat y Diversidad. Grafikar, La Plata. Castellanos, A. 1965. Estudio fisiográfico de la Provincia de Corrientes. Instituto de Fisiografía y Geología.Universidad Nacional del Litoral. CECOAL (Centro de Ecología Aplicada del Litoral). 1981. Estudios Ecológicos en el área del embalse Yaciretá. EBY, Corrientes. Contreras, F.I., E.A. Ojeda, and S. Contreras. 2014. Aplicación de la línea de costa en el estudio morfométrico de las lagunas de las lomadas arenosas de Corrientes, Argentina. Contribuciones Científicas GÆA: 26: 65-78. Contreras, F.I., J. Meza Scipioni, N. Hernández, and F.J. Ruiz-Díaz. 2017. Cambios morfométricos de lagunas aluviales del río Paraná y su incidencia en la diversidad íctica. Revista Veterinaria: 2: 51-55. Corrales, M.A. 1979. Contribución al conocimiento del zooplancton del Alto Paraná. Ecosur 6: 185-205. Corrales de Jacobo, M.A and S.M. Frutos. 1982. Zooplancton: composición structural y densidad de poblacion. Fluctuaciones con los ciclos

Herbst, R. and J. Santa Cruz. 1999. Mapa litroestratigráfico de la provincia de Corrientes. D´Orbygniana 2:1-69. Iriondo, M.H., 1991. El Holoceno en el Litoral. Comunicaciones Museo Provincial de Ciencias Naturales “F. Ameghino” (Nueva Seire), 3 (1): 1-40.
 Iriondo, M. 2004. Large wetlands of South America: a model for Quaternary humid environments. Quaternary International 114: 3–9. Iwaszkiw, J.M., F. Firpo Lacoste, and A. Jacobo. 2010. Relevamiento de la ictiofauna de la laguna Camba Cué, isla Apipé Grande, Corrientes, Argentina. Revista del Museo Argentino de Ciencias Naturales 1: 81-90. Junk, W.J. (ed.). 1997. The Central Amazon Floodplain: Ecology of a Pulsing System. Ecological Studies. Springer-Verlag, Berlin Heidelberg. Junk, W.J, P.B. Bayley, and R.E. Sparks. 1989. The flood pulse concept in river– floodplain systems. Canadian Special Publications for Fisheries and Aquatic Sciences 106: 110–127. Kopuchian, C., L. Campagna, D.A. Lijtmaer, G.S. Cabane, N.C. García, P.D. Lavinia, P.L. Tubaro, I. Lovette, and A.S. Di Giacomo. 2020. A test of the riverine barrier hypothesis in the largest subtropical river basin in the Neotropics. Molecular Ecology 2020:1–13. https://doi.org/10.1111/

mec.15384 Lopretto, E.C. 1995. Crustacea Eumalacostraca, In: E.C. Lopretto and G. Tell (eds.). Ecosistemas de aguas continentales. Ediciones Sur, La Plata. pp. 1001–1039. McClain, M.E and R.J. Naiman. 2008. Andean influences on the biogeochemistry and ecology of the Amazon River. Bioscience 58: 325–338. Melack, J.M. and B. Forsberg. 2001. Biogeochemistry of Amazon floodplain lakes and associated wetlands. In: M.E. McClain, R.L. Victoria and J.E. Richey (eds.). The Biogeochemistry of the Amazon Basin and its Role in a Changing World. Oxford University Press, Oxford. pp. 235-276. Mitsch, W.J. and E. Jørgensen. 2003. Ecological engineering: A field whose time has come Ecological Engineering 20: 363–377. Wetland Science & Practice October 2020 281


Morton, L.S. 2004. Moluscos fósiles de agua dulce de la Formación Ituzaingó, Plioceno de Corrientes. In: F.G. Aceñolaza (ed.). Temas de la Biodiversidad del Litoral Fluvial Argentino. Instituto Superior de Correlación Geológica, Universidad Nacional de Tucumán, Miscelánea. 12: pp. 45-48. Neiff, J.J. 1978. Fluctuaciones de la vegetación acuática en lagunas del valle del río Paraná en la transección Paraná-Santa Fe, entre 1970 y 1977. Physis 38: 41-53. Neiff, J.J. 1986. Aquatic Plants of the Paraná System. In: B.R. Davies. and K.F. Walker (eds.). The Ecology of River Systems. Dr. W. Junk Publishers, The Netherlands. pp. 557-571. Neiff, J.J. 1990. Aspects of primary productivity in the lower Paraná and Paraguay riverine system. Acta Limnologica Brasiliensia 3: 77-113. Neiff, J.J. 2003. Distribución de la vegetación acuática y palustre del Iberá. In: A. Poi (ed.). Limnología del Iberá. Aspectos físicos, químicos y biológicos de sus aguas. Editorial de la Universidad Nacional del Nordeste, Corrientes. pp. 17-65. Neiff, J.J. 2004. El Iberá...en peligro? Fundación Vida Silvestre de Argentina, Buenos Aires. Neiff, J.J. and S.L. Casco 2017. Lluvias y sequías: los cambios históricos de la vegetación. In: A.S.G Poi (comp.). Biodiversidad en las aguas del Iberá. Editorial de la Universidad Nacional del Nordeste, Corrientes. pp 41-74. Neiff, J.J., S.L. Casco, A. Cózar Cabañas, A. Poi and B. Úbeda Sánchez. 2011. Vegetation diversity in a large Neotropical wetland during two different climatic scenarios. Biodiversity and Conservation 20: 2007-2025. Neiff, J.J. and M. Neiff. 2013. Evaluación de los impactos del cambio climático sobre el ecosistema natural y la biodiversidad. Esteros del Iberá (Argentina). CEPAL-Serie Medio Ambiente y Desarrollo, Chile: Naciones Unidas.
 Neiff, J.J., A. Poi de Neiff and M.B. Canón Verón. 2009. The role of vegetated areas on fish assemblage of the Paraná River floodplain: effects of different hydrological conditions. Neotropical Icthyology 7: 39-48. Odum, H.T. and B. Odum. 2003. Concepts and methods of ecological engineering. Ecological Engineering 20: 339–361. Orfeo, O. and J.C. Stevaux. 2002. Hydraulic and morphologic characteristics of middle and upper reaches of the Paraná River (Argentina and Brazil). Geomorphology 44: 309-322. Orfeo, O. 2005. Historia geológica del Iberá, como escenario de biodiversidad. Miscelanea 14: 71-78. Orfeo, O. and J.J. Neiff. 2008. Esteros del Iberá. In: Comisión Sitios de Interés Geológicos de Argentina (ed). Sitios de interés Geológico de la República Argentina. Instituto de Geología y Recursos Minerales, Servicio Geológico Minero Argentino, Anales, Buenos Aires. pp. 415-425.

Poi, A.S.G. (comp.). 2017. Biodiversidad en las aguas del Iberá. Editorial de la Universidad Nacional del Nordeste, Corrientes. Poi, A. S.G., J.J. Neiff, S. L. Casco, B. Úbeda Sánchez, and A. Cózar Cabañas. 2017. El agua de los esteros, lagunas y ríos. In: A.S.G. Poi (comp.). Biodiversidad en las Aguas del Iberá. Editorial de la Universidad Nacional del Nordeste. Corrientes. pp. 21-39. Popolizio, E. 1977. Contribución a la Geomorfología de la Provincia de Corrientes (Argentina). Geociencias VII: 1-45. Popolizio, E. 1981. Geomorfología del Macrosistema Iberá. Informe ICA, Corrientes, 186 p Popolizio, E. 2004. El Paraná, un río y su historia geomorfológica. Tomos I y II. (PhD thesis). Centro de Geociencias Aplicadas. Resistencia. Thoms, M.C. 2003. Floodplain-river ecosystems: lateral connections and the implications of human interference. Geomorphology 56: 335–349. Tiner, R.W. 2017. Wetland Indicators: A Guide to Wetland Formation, Identification, Delineation, Classification, and Mapping. Lewis Publ., Boca Raton, Florida. Tiner, R.W. 2003. Geographically isolated wetlands of the United States. Wetlands 23: 494-516. Tockner K., F. Schiemer, and J.V. Ward. 1998. Conservation by restoration: the management concept for a river- floodplain system on the Danube River in Austria. Aquatic Conservation: Marine and Freshwater Ecosystems 8: 71–86. Úbeda B., A.S. Di Giacomo, J.J. Neiff, S.A. Loiselle, A.S.G. Poi, J.A. Gálvez, S.L. Casco, and A. Cózar. 2013. Potential effects of climate change on the water level, flora and macro-fauna of a large neotropical wetland. PLoS ONE 8: e67787. https://doi.org/10.1371/journal. pone.0067787. Vannote R.L., G.W. Minshall., K.W. Cummins, J.R. Sedell, and C.E. Cushing. 1980. The river continuum concept. Canadian Journal of Fisheries and Aquatic Sciences 37: 130–137. Varela, M.E. and J.A. Bechara. 1981. Bentos. In: Convenio ICA-CECOAL Investigaciones ecológicas en el macrosistema Iberá. Informe final. Corrientes. pp. 75-85. Varela, M.E., J.A. Bechara, and N.L. Andreani. 1983. Introducción al estudio del bentos del Alto Paraná. Ecosur 10: 103-126. Ward, J.V. and J.A. Stanford. 1995. The serial discontinuity concept: extending the model to floodplain rivers. Regulated Rivers 10: 159–168. Wiens, J.A. 1989. Spatial scaling in ecology. Functional Ecology 3: 385–397. Wiens, J.A. 2002. Riverine landscapes: taking landscape ecology into the water. Freshwater Biology 47: 501–515.

Orfeo, O., F. Colombo, and J.J. Neiff. 2014. Desplazamientos laterales del curso del río Paraná en las cercanías de la ciudad de Corrientes. XIV Reunión Argentina de sedimentología: 206-207.

Wiens, J.A. 2009. Landscape ecology as a foundation for sustainable conservation. Landscape Ecol. 24:1053–1065. https://doi.org/10.1007/ s10980-008-9284-x.

Pacella, L.F. and M. Di Pasquo. 2020. Aportes de la palinología al conocimiento de los embalsados holocenos de corrientes, argentina. Rev. Brasileira de Paleontología 23 (1): 63–72,

Zalocar de Domitrovic, Y. 1990. Efecto de las fluctuaciones del nivel hidrométrico sobre el fitoplancton en tres lagunas isleñas en el área de la confluencia de los ríos Paraná y Paraguay. Ecosur 16: 13-29.

Paggi, J. and S. José de Paggi. 1974. Primeros estudios sobre el zooplancton de las aguas lóticas del Paraná Medio. Physis 33: 91-114

Zalocar de Domitrovic, Y. 1992. Fitoplancton de ambientes inundables del río Paraná (Argentina). Revista de Hydrobiologia Tropical 25: 177-188.

Poi de Neiff, A. 1981. Mesofauna relacionada a la vegetación acuática en una laguna del valle del Alto Paraná. Ecosur 8: 41-53.

Zalocar de Domitrovic, Y. 2003. Fitoplancton de lagunas y cursos de agua del Sistema Iberá. In: A. Poi de Neiff (ed.) Limnología del Iberá. Aspectos físicos, químicos y biológicos de las aguas. Editorial de la Universidad Nacional del Nordeste, Corrientes. pp. 85-127.

Poi de Neiff, A. (ed.). 2003. Limnología del Iberá. Aspectos físicos, químicos y biológicos de las aguas. Editorial de la Universidad Nacional del Nordeste. Sigma, Buenos Aires. Poi de Neiff, A., J.J. Neiff, and S.L. Casco. 2006. Leaf litter decomposition in three wetlands of the Paraná River floodplain. Wetlands 26: 558-566. 282 Wetland Science & Practice October 2020

Zalocar de Domitrovic, Y., A. Poi de Neiff, and S.L. Casco. 2007. Abundance and diversity of phytoplankton in the Paraná River (Argentina) 220 km downstream of the Yacyretá reservoir. Brazilian Journal of Biology 67: 53-63. 10.1590/S1519-69842007000100008.


WETLAND RESEARCH

Urban Wetland Trends in Three Latin American Cities during the Latest Decades (2002-2019): Concón (Chile), Barranquilla (Colombia), and Lima (Peru) Carolina Rojas1,2, Juanita Aldana-Domínguez3, Juan Munizaga4, Paola Moschella5, Carolina Martínez6, and Caroline Stamm7

ABSTRACT etlands are valuable but threatened natural resources worldwide. While providing a wealth of environmental benefits, wetlands play a vital role in temporarily storing flood waters and thereby reducing the risk of damaging floods. This is important given the predicted impacts of climate change, especially along the world’s coastline and coastal cities. The continued expansion of urban areas is posing a risk to wetlands in and around metropolitan areas. In this article we examine wetland trends in urban areas in three Latin American countries – Chile, Colombia, and Peru.

W

INTRODUCTION Wetlands, including their associated vegetation and the water bodies, cover at least 10% of the planet (Davidson et al. 2018). They are disappearing, even though they are relevant ecosystems for ecological balance and for mitigating the effects of climate change. It is estimated that since 1900, the world has lost more than 50% of these ecosystems, and the increase in urbanization has been identified as one of the main causes for this loss (Boyer and Polasky 2004; Faulkner 2004; Bishop et al. 2006; González et al. 2014). In fact, more than 55% of the world’s population lives in cities, and it is expected that by 2050, this figure will reach 68% (United Nations 2018). As urban growth increases, wetland area decreases. Unfortunately, Latin America leads this worldwide tendency, reporting a loss of 59% of wetlands over the last decades (1970-2015) (Darrah et al. 2019). This loss is combined with the fact that it is one of the poorest and most economically unequal regions on the planet (Cepal 2019). Wetlands are particularly relevant for cities because: 1) most large cities are located on the coast, 2) world urban growth has been concentrated in low-elevation coastal 1 Institute of Urban Studies, Pontificia Universidad Católica de Chile, Santiago, Chile. 2 Correspondence author contact: carolina.rojas@uc.cl. 3 Department of Chemistry and Biology, Universidad del Norte Colombia, Barranquilla, Colombia. 4 EULA Center, Universidad de Concepción, Concepción, Chile. 5 Department of Humanities - Geography and Environment Section, Pontificia Universidad Católica del Perú, Lima, Perú. 6 Department of Geography, Pontificia Universidad Católica de Chile, Santiago, Chile. 7 Institute of Urban Studies, Pontificia Universidad Católica de Chile, Santiago, Chile.

zones, and 3) the assistance of wetlands when facing frequent urban disasters has not been considered. It is a fact that wetlands mitigate the impact of flooding, which helps make cities more resilient (Ramsar 2019). In relation to this, according to the United Nations (2014), 233 of the world’s cities are located in zones that are high at risk for flooding, affecting approximately 663 million people. These urban areas require precaution and more resilient infrastructure. The importance of wetlands could be even greater when facing extreme events, such as the flooding that occurred in Phoenix, USA in 2014 (Kim et al. 2017). In fact, in the northeastern USA, wetlands helped save $625 million dollars in direct damage from the floods caused by Hurricane Sandy in 2012 (Narayan et al. 2017). One of the main ecosystem services of wetlands is protecting the coast. They are also important for cities, as they act as carbon sequestering systems, purify water, and maintain biodiversity and ecological processes. They even provide a place for recreation and relaxation for urban residents (Maltby and Acreman 2011; McInnes 2014). Without a doubt, the loss of wetlands is affecting the sustainability and resilience of our Latin American cities. This is especially true for coastal cities that face the challenge of being prepared for the increase in frequency of natural disasters, such as urban flooding. Ramsar (2018) also highlights the role of wetlands in cities by decreasing the impact of urban flooding caused by strong rainfall, as well as providing a buffer from swells and tsunamis. However, Latin America is dealing with a loss of wetlands, despite being one of the regions most exposed to flooding (UN 2011). Evidently, this tendency also increases the region’s vulnerability to climate change (Seto et al. 2011; Hallegatte et al. 2013). We are facing an alarming scenario, where sustained urban growth is the generalized trend in Latin American cities (UN 2018) which in many cases occurs at the expense of natural spaces (Rojas et al. 2013; Aldana-Domínguez et al. 2019). Facing pressure to expand, cities have converted wetlands to developable land by legal and illegal landfills as is the case in Chile and Argentina (Rojas 2018; Pintos and Sgroi 2012). Legal clearing is protected by urban laws that do not incorporate factors such as ecological connectivity, integration with the coast, or geomorphological Wetland Science & Practice October 2020 283


TABLE 1. Selected images per wetland study area.

Study Area Aconcagua Ciénaga de Mallorquín Pantanos de Villa

Collection Image Source Year 2004 – 2019 Google Earth Pro – ArcGIS Pro 2002 – 2019 Google Earth Pro – ArcGIS Pro 2002 – 2019 Google Earth Pro

FIGURE 1: Location of the studied wetlands.

284 Wetland Science & Practice October 2020

Error (RMS) 0.45 – 0.32 0.69 – 0.62 0.98 – 0.73

processes (Rojas et al. 2019). As for the illegal clearing, in addition to the environmental impact, high risk situations are generated, caused by ground instability. This progressive growth of Latin American cities has altered and fragmented wetlands, whether it be for housing or transportation infrastructure, creating coastal zones with degraded and devalued ecosystems that are marginalized from the territory. Because of this situation in Latin America, the study of urban wetlands is highly relevant,


as it can contribute to our base of knowledge and support conservation initiatives. Urban wetlands have been little studied and defined, to the point that the Ramsar Convention (Secretary of the Ramsar Convention) or global treaty for the conservation of wetlands recognizes this debt, admitting that urban wetlands have been forgotten (Hettiarachchi et al. 2015). Because of this, these authorities have highlighted the relevance of urban wetlands, and in 2018, they declared that they were key sites for making cities healthy and habitable. On the other hand, the lack of information on the boundaries of urban wetlands has further complicated their recognition in the planning of cities. They are in fact a unique ecosystem. It is very challenging to physically define them with remote sensing techniques and geographic information systems, especially when they coexist in a heterogeneous landscape and are located in coastal cities (Adam et al. 2014; Gibril et al. 2020). Furthermore, they do not have a uniform plant coverage, are highly dynamic and their spectral reflectance is easily disrupted. Additionally, land use has further increased their variability, causing modifications in their vegetation and water levels (Gallant 2015). The objective of this study is to carry out a spatial and multi-temporal analysis with remote sensors, to determine wetland surfaces and changes in three cities in Chile, Colombia and Peru. It seeks to analyze changes in land use that have led to loss and gain in urban wetlands during the period (2002-2019) when the greatest loss of these ecosystems has been reported (Darrah et al. 2019). Land changes are analyzed with high-resolution satellite images, which offset problems of definition by other sensors such as Landsat, as they have only been available for the last two decades in the wetlands situated in the Latin American coastal cities of Barranquilla, Lima, and Concón. These data will allow for discovering the main spatial dynamics related to the reduction and alteration of these ecosystems, thereby taking a first step towards the recognition of the status of these ecosystems. The chosen cities are part of the “Urban Wetlands in Latin America: a solution for sustainable cities SDG 11” Project (2019 -2021); cities that have experienced rapid urbanization and are vulnerable to climate change.

TABLE 2. Land use covers, definitions and reference signatures.

OVERVIEW OF URBAN WETLANDS IN LATIN AMERICA Although Latin American countries have joined the Ramsar Convention (1981), and represent 11% of the world’s wetlands, the region leads in wetland loss, reaching 59% (Darrah et al. 2019). Specifically, the region has observed losses of wetland surface in cities in Chile, Peru, Colombia, Argentina and Brazil. Decreases are registered on the Brazilian coast (Sousa et al. 2011; Wittmann et al. 2015), the Andes and Caribbean regions of Colombia (Patino Wetland Science & Practice October 2020 285


areas. The pressure to build has led to the elimination of wetlands in cities spanning from north to south, especially in the southern-central zone, which happens to have a high recurrence of waterlogging and flooding (Rojas et al. 2014). The Aconcagua-Concón wetland is under pressure from industrial activity and from urban growth. There is evidence of an expansion in the real estate market for second homes (Hidalgo et al. 2016; Martínez et al. 2020), and a concentration of economic activities (ports, tourism and services). In addition, since 2015 this coastal zone has been seriously affected by extreme events associated with climate change, including swells, meteotsunamis, coastal erosion and tsunamis like Japan 2011 and Illapel, Chile (Martínez et al. 2011; Martínez et al. 2018; Carvajal et al. 2017, Campos-Cava 2016); all of which have caused considerable damage to the infrastructure and connectivity (Winckler et al. 2017). Additionally, the areas adjacent to the wetland (the Aconcagua River estuary) have been categorized as a low-quality landscape because of the high degree of anthropic intervention, which has led to a further loss of naturalness on the coastal landscape (Rangel-Buitrago et al. 2018). Colombia TABLE 3. Land uses changes in wetlands In Colombia, wetlands make up more than 30 million ha, Wetlands Total Total which is equivalent to 26% of its territory. Wetlands have historically been associated with the development of human Losses Gains cultures (Jaramillo et al. 2015). The main cause of wetland Aconcagua 33 10 transformation has been the change in land use to pastures La Ciénaga de Mallorquín 224 81 for raising cattle, agriculture and deforestation, and to a Pantanos de Villa 11 2 lesser extent, urbanization (Patino and Estupinan-Suarez Total 268 93 2016). These anthropic changes have led to the transformation of 24% of the country’s FIGURE 2. Distribution of the land uses and cover types detected in the studied wetlands and surrounding wetlands. It is estimated that areas for the Aconcagua study area (Concón, Chile) from 2004 to 2019. by 2025, the main factors of change in wetlands will be, in first place, the expansion of ranching, and in second place, urban and transportation development (Ricaurte et al. 2018). The Magdalena-Cauca basin (i.e., where the Metropolitan Area of Barranquilla is located) is expected to suffer the greatest changes in land use, with negative effects on wetlands. Barranquilla is a coastal city located on the Magdalena River estuary in the Caribbean Sea. This estuary creates a vast area of wetlands that have been greatly altered by anthropic and Estupinan-Suarez 2016) and the Luján River basin in Argentina (Pintos y Sgroi 2012). However, figures of each Latin American country’s contribution and the distribution of most of the Latin American wetlands are still unknown, which is worrisome (Mitsch and Gosselink 2015). Coastal cities in Latin America are vulnerable to the effects of climate change, which together with the socionatural risks like flooding, means that more people are exposed to danger. Chile, Colombia and Peru have been chosen as example countries with wetlands that are subject to urbanization and vulnerability to climate change. The three countries have experienced significant urban growth and project further increases for 2030. Additionally, 80% of their population lives in cities, which are representative of the environmental conflicts caused by urbanization and conflicts with wetlands located in cities. Chile In Chile, more than 80% of the population lives in cities. There are more than 4 million ha of wetlands, but only 3% is protected. None of these protected wetlands are in urban

286 Wetland Science & Practice October 2020


activities throughout history and is highly vulnerable to the effects of global change (Aldana-Domínguez et al. 2018; Rodríguez 2015). The transformation of these ecosystems has had a negative impact on the ecosystem services, mainly on the regulation services (Aldana-Domínguez et al. 2019), which means that new guidelines that allow for recovering and conserving the urban ecosystems must be a priority. Peru In Peru, the expansion of cities is one of the main causes of wetland degradation (MINAM 2015), especially in the coastal and desert strip where wetlands have a special value as nuclei of biodiversity and freshwater reserves. The wetland ecosystems cover a total surface of 6.9 million ha, which is equivalent to 5.4% of the territory (MINAM 2019). Among them, the coastal wetlands only occupy 0.04%, and are especially vulnerable to growing real estate pressure and the impact of urbanization on the aquifer. For example, the expansion of the city of Lima between 1990 and 2013 led to the loss of 203 ha of wetlands (Moschella 2018). Studies on the main wetlands in Lima show the degradation and loss of ecosystem services caused by anthropic pressure (Aponte and Cano 2013; Moschella 2012; Pulido and Bermúdez 2018). Although there is a legal framework for protecting wetlands, there are weaknesses in the instruments of protection and an adequate regulation for the application of norms is lacking (Ramírez Aponte 2018). Further studies are also required to understand the hydrology of these ecosystems and contribute to their conservation (Rodríguez 2017). In this sense, the National Wetland Strategy (MINAM 2015) has identified a lack of studies on the valuing and management of wetlands as well as a weak participation for the conservation of these ecosystems. METHODOLOGY Study Areas The following urban coastal wetlands were chosen for this study: 1) Aconcagua (also known as the Concón Wetland) in the city of Concón and associated with the Andean Aconcagua River in the Valparaíso Region (Chile), 2) la Ciénaga de Mallorquín in Barranquilla (Colombia), and 3) Pantanos de Villa in Lima (Peru). They are all located in coastal areas and are subject to the different pressures of changes in use from urbanization (Figure 1). Data Processing Free satellite images from the last two decades (2002-2019) were selected from the collection available from the Quickbird satellite on Google Earth Pro. Then, a buffer of approximately 500 meters from the perimeter of each wetland was defined. The images were georeferenced and classified by photointerpretation at a scale of 1:2.000 in ArcGIS Pro.

TABLE 4. Land cover changes in Aconcagua study area from 2004-2019, the analyzed image included the wetland and its surrounding areas. (Note: Any difference in net change totals is due to computer round-off.)

Land use Water bodies Roads Dunes Wetlands Other vegetation Plantation forest Beaches Bare soils Urban Agriculture Grasslands Industrial

Aconcagua between 2004 and 2019 Losses Gains Net Change (ha) (ha) (ha) 0 0 -5 -33 -6 -9 -10 -64 -1 0 -29 0

9 0 0 10 4 1 5 12 15 3 66 31

8 0 -5 -23 -2 -7 -5 -52 14 3 37 31

TABLE 5. Land cover changes in Ciénaga de Mallorquín study area from 2002-2019, the analyzed image included the wetland and its surrounding areas. (Note: Any difference in net change totals is due to computer round-off.)

Land use Water bodies Roads Other vegetation Beaches Bare soils Urban Grasslands Permanent wetland Semi-permanent wetland Industrial

Ciénaga de Mallorquín between 2002 and 2019 Losses Gains Net Change (ha) (ha) (ha) -14 0 -77 -20 -82 -1 0 -175

128 7 18 36 39 70 17 3

114 6 -59 15 -42 70 17 -172

-49 -6

78 28

29 22

Wetland Science & Practice October 2020 287


FIGURE 3. Gains and losses in the Aconcagua Wetland (Concón, Chile) from 2004 to 2019.

Preparation and Selection of Satellite Images for Land Classification Satellite images were selected from Google Earth Pro, according to the following criteria: • Time: Search of the collections that represent changes over the last 20 years. • Cloudiness: Determining factor for the selection of an image in its next classification as cloudiness presents one of the main problems for detection. • Seasonality: Preference for images with similar seasonal conditions (winter – summer).

These criteria defined the studied time horizon from 2002 to 2019, and the 2.5-meter resolution Quickbird sensor (See Table 1). A priori these images also strengthen the detection of smaller surfaces like urban wetlands. The QuickBird sensor images were georeferenced by ArcGIS Pro by taking control points in FIGURE 4. Distribution of the land uses and cover types detected in the studied wetland and KML format on Google Earth Pro. A total of surrounding area for the Ciénaga de Mallorquín study area (Barranquilla, Colombia) from 18 control points was taken in each image for 2002 to 2019. each wetland, which were complemented with points on the ArcGIS Pro basemap and field work in March 2020. Once georeferencing was completed, points with an error greater than 1 meter were eliminated (Table 1). Classification and Land Use Covers Classification was done through a combination of photointerpretation and image tracing of the Quickbird images on ArcGIS Pro and the calculation of the Normalized Difference Vegetation Index (NDVI) that allowed for observing the different types of vegetation and separating the bodies of water and the uncovered land. The recognition of the categories was based on the definitions and reference images shown in Table 2 that allowed for distinFIGURE 5. Gains and losses in the permanent wetland (a) and the semi-permanent guishing natural areas from artificial areas and wetland (b) at the Ciénaga de Mallorquín (Barranquilla, Colombia) from 2002 to 2019. using the time series of imagery to observe The figures also show changes in wetland type (permanent v. semi-permanent). The conversion towards water bodies accounted for most of the permanent wetland loss, with trends from 2002/2004=2019. Surroundchange to semi-permanent wetland also a major factor in loss of this type. ing vegetation that is not part of the wetland was not classified in detail. For the Ciénaga de Mallorquín, two types of wetlands were identified following the Colombian wetland classification system (Ricaurte el al. 2019): permanent wetland (permanently flooded), and semi-permanent wetland (periodically flooded; mainly covered by mangrove forests in this area).

288 Wetland Science & Practice October 2020


Validation of Land Covers The validation of coverages was done through the capturing of 840 points of control, whose verification was optimized in GIS and with field work conducted in March of 2020. Then, the veracity of the interpreted classification was evaluated with the Kappa Statistic Index.

the total loss of 33 ha were lost from 2004 to 2019, mainly caused by the conversion to grasslands (17 ha), followed by bare soils (8 ha), beaches (3 ha) and finishing with less pressure of urban zones (2 ha) (Figure 3). The losses were somewhat compensated for by a gain of 10 ha from bare soils (9 ha) and beaches (1 ha). Besides the loss of wetlands, bare soils experienced a loss of 64 ha and gain of 12 ha for a net loss of 52 ha from 2004-2019. This loss accounted for some of the increase in grasslands surrounding the wetland. Some of the new grasslands also came from beaches. A similar situation has occurred with the dunes – conversion of dunes to grassland; they experienced a net loss of 5 ha (Table 4).

RESULTS Land Covers The results of the land cover typing for the three wetland areas are shown in Figures 2, 4, and 6; they generated kappa index values of >85% of precision. Therefore, the maps allow for making a statistically valid interpretation. The distribution and changes in land use are reported in Tables 3-5, where values from the initial year, gains, losses TABLE 6. Land cover changes in Pantanos de Villa study area from 2002and net changes can be observed in the categories identified 2019, the analyzed image included the wetland and its surrounding areas. for each wetland. (Note: Any difference in net change totals is due to computer round-off.) A sum of 268 ha of decrease was observed totalizing Pantanos between 2002 and 2019 the three studied wetlands, of which 175 correspond to Land use Losses Gains Net Change loss (Losses – Gains). The main cause was an increase in (ha) (ha) (ha) artificial and productive covers such as urbanization and Water bodies 0 9 9 grasslands, among others. Urbanization increased by a total of 162 ha considering the three studied cases. The land use Roads -1 5 4 covers changes observed over the last two decades themWetlands -11 2 -8 selves confirm the general trend of wetland loss in the Latin Beaches -11 2 -9 American region reported by Darrah et al. (2019). Bare soils -76 7 -70 Of the studied wetlands, La Ciénaga de Mallorquín Urban -8 86 78 (Colombia) has without a doubt suffered the greatest surface loss, followed by Aconcagua (Chile) and Pantanos Agriculture -1 0 -1 de Villa (Perú). Additionally, among the three wetlands, Green areas -3 0 -3 Pantanos de Villa is under the most pressure from urbanization in its FIGURE 6. Distribution of the land uses and cover types detected in the studied wetland and sursurroundings (i.e., it is virtually surrounding area for the Pantanos de Villa wetland (Lima, Peru) from 2002 to 2019. The larger “Green rounded by urban development with Area” is a private golf course. the exception of the “Green Area”). Aconcagua Wetland Area (Concón, Chile) In the case of the Aconcagua wetland study area, during the period of analysis the main interactions and exchanges were between the wetland, grassland, bare soils, beaches and urban and industrial areas (Figure 2). First, the category experiencing the greatest gain was grasslands, with 37 ha of net change, closely followed by industrial area with 31 ha and then by urbanization with a net gain of 14 ha (Table 3). The Aconcagua wetland itself experienced a net loss of 23 ha, but with

Wetland Science & Practice October 2020 289


FIGURE 7. Gains and losses in the Pantanos de Villa Wetland (Lima, Perú) from 2002 to 2019. There were small changes in wetland area, mainly a few hectares of losses.

La Ciénaga de Mallorquín Wetland Area (Barranquilla, Colombia) Both permanent and semi-permanent wetlands at La Ciénaga de Mallorquín changed during the study period (Table 5 and Figures 4 and 5). The main wetland change was the loss of permanent wetland surface (net loss of 172 ha). This loss was mainly due to coastal erosion resulting from the displacement of the sand bar that separates the wetland from the Caribbean Sea (Figure 5a). This contributed 94 ha to the overall net increase in the water bodies coverage. The water bodies are represented by the ocean and the river. The wetlands have connections to both, but are separated by a sand bar and a spur. The retraction of the sand bar was documented by Rivillas-Ospina and others (2018) who noted an erosion rate of 0.14 m/year from 1973 to 2016. The permanent wetland was also reduced by an increase in the semi-permanent wetland (52 ha), probably due to the natural formation of mangrove areas on the western side of the Ciénaga de Mallorquín and a recent mangrove reforestation effort. The formation of beaches and conversion to urban development were also responsible for the loss of 22 ha and 6.7 ha, respectively, of permanent wetland. On the other hand, the semi-permanent wetland gained 52 ha at the expense of permanent wetland and 25 ha from bare soils, while it lost some surface area mainly to water bodies (12 ha), bare soils (11 ha), grasslands (10 ha), and urban zones (8 ha; Figure 5b). Although the Ciénaga de Mallorquin is included in the Ramsar Site “Estuarine system of the Magdalena River Ciénaga Grande de Santa Marta”, anthropic pressures continue to affect it.

290 Wetland Science & Practice October 2020

Another significant change in the area surrounding the wetland was the loss of other vegetation (59 ha; Table 5), which included the last few fragments of tropical dry forest in the area. The tropical dry forest is one of the most threatened ecosystems and at risk of collapse in Colombia (Etter et al. 2017), and it is also a key ecosystem for the supply of ecosystem services in Barranquilla (Aldana-Domínguez et al. 2019). It continues to lose surface area, mainly because of urbanization. For the entire study area, as expected, urbanization has increased (70 ha; Table 5), affecting both wetlands and tropical dry forest. Pantanos de Villa Study Area (Lima, Peru) For the Pantanos de Villa study area, the greatest change from 2002 to 2019 was the increase of urban areas (86 ha; Table 6 and Figure 6), which mostly happened on bare soils, and to a lesser extent on wetlands and beaches. Wetland coverage declined by 11 ha: 6 ha due to urbanization, 2 ha lost to bare soils, and the remainder to beaches and roads. Wetland gained only 2 ha, with the most significant corresponding to wetland expansion over an unbeaten path (Figure 7). Although the variation in the wetland’s extent seems limited, it is worth mentioning that this wetland has been converted to urban land for several decades before the period of this analysis. While most of the new urban areas are registered here as conversions from bare soils without use, these bare soils are actually cleared wetland (i.e., they were mainly filled to dry the land for future plotting and building). Currently, the wetland is practically restricted to the area that is protected and recognized as a Ramsar site. Since its borders are mostly developed, it is most likely that its expanse will not vary. Additionally, it is worth mentioning that there are significant differences between the degree of urban planning and its impact on the wetland’s functioning. To the northeast of the wetland, irregular low-income settlements and filled wetlands can be found. Here, recent land-use changes affect the main springs that supply the wetland and the water channels that connect them, putting the quality and quantity of surface flow entering the wetland at risk. Specifically, there is an increase of dwellers that use the spring water as they lack piped drinking water. Also, a near and unstable landfilling might block the channels. For instance, a public sports field obstructs the water flow. Meanwhile, the southwest end presents mostly beach urbanization through construction of high-income condominiums, which have drained the wetlands and have also reduced the ecosystem’s connection with the coast.


CONCLUSION The analysis of Quickbird images from the period of 2002-2019 has allowed for the creation of valid maps that show land use cover, including wetlands during two time periods. Consequently, it has permitted spatial and temporal tracking of wetland trends in selected urban wetlands in Latin America and provided information on land cover changes over the past 20 years. Our study shows important reduction in surface area of most of the studied urban wetlands, mainly caused by three factors: coastal erosion/creation of water bodies, expansion of grassland and growth of urban areas surrounding them. The Ciénaga de Mallorquín wetland has suffered the greatest loss, caused by coastal dynamics and changes in the water body and vegetation. Coastal erosion is a complex phenomenon and, in this area, it is related with oceanographic processes (i.e., currents and waves generated by the wind and tides) and the anthropogenic impacts originated by the construction of infrastructures to maintain the Barranquilla port and the navigable channel (Rivillas-Ospina et al. 2018). The spreading of grasslands could be interpreted as pre-construction activity, where grassland increase over bare soils and agricultural areas mostly in the case of Aconcagua. The Aconcagua wetland area, although located in an area under a great deal of historic pressure from industrial activity, is currently being altered by an increase of grasslands and bare soils specially when the dunes started to increase with low vegetation due to natural conditions as a precipitation, which is also affecting the losses in beaches. The Aconcagua wetland is lesser affected by urban growth and agriculture areas. Clearly, urbanization is affecting these three areas. The Ciénaga de Mallorquin is the most urbanized wetland studied, losing 15 ha due to urban expansion. Two sectors of urban growth are recognized: the eastern side of the wetland where unplanned settlements built by displaced people from other Colombian regions, and even from other countries, come to Barranquilla seeking better life opportunities. And on the western side of the wetland, in addition to the informal settlements, there are the planned social interest urban developments to fulfil the housing deficit of the population with lower monetary resources. In a similar way, in Pantanos de Villa and its surrounding areas the urban areas have increased at a greater speed because of pressure for two type of development: 1), informal settlements or housing units constructed illegally (i.e., unplanned and with scarcity of services such as drinking water) and 2) beach condominiums or planned settle-

ments, which were built on cleared wetlands and the area surrounding the protected zone (i.e., designated Ramsar Site). The urbanization of Aconcagua wetland is lesser than grassland growth and, in this period, impacted due to urbanization process was produced before eighty decades. Overall, the urban wetlands have lost surface, confirming the Latin American trend. Urban expansion, mainly for housing, is the major impact but urbanization may also be responsible for a change in beaches and grassland. Pressures surrounding the wetlands that could be interpreted as an urbanization transformation process were also identified e.g., kind of settlements in Pantanos de Villa (Perú). Latin America faces a great challenge in advancing towards urban sustainability according to the 17 Sustainable Development Goals (SDG established by the 2030 Agenda (UN, 2015) where the Goal 11 have a target to make cities and human settlements inclusive, safe, resilient and sustainable. The region also needs to improve urban planning and management regarding the identification, definition, and norms for permitted uses of urban wetlands and their surroundings. Urban wetlands should be “recognized spaces” in the city – natural habitats important for the well-being of the city residents and kept safe, resilient and sustainable facing the challenges of climate change. n ACKNOWLEDGEMENTS This study is part of the research project “Urban Wetlands in Latin America: a solution for sustainable cities SDG 11” (2019 -2020). It was financed by the CODS – Latin American Sustainable Development Goals Center, whose general objective is to strengthen interdisciplinary research and cooperation on urban wetlands, recognizing that they are key ecosystems fundamental for fostering more sustainable and resilient cities (i.e., by 2030, as mandated in the SDG 11 for Latin American countries), and therefore strengthening their connection with public policy. REFERENCES

Adam, E., O. Mutanga, J. Odindi, and E.M. Abdel-Rahman. 2014. Land-use/cover classification in a heterogeneous coastal landscape using RapidEye imagery: evaluating the performance of random forest and support vector machines classifiers. International Journal of Remote Sensing 35(10): 3440-3458. Aldana-Domínguez, J., C. Montes, and J.A. González. 2018. Understanding the past to envision a sustainable future: A social-ecological history of the Barranquilla Metropolitan Area (Colombia). Sustainability 10: 1-18. Aldana-Domínguez, J., I. Palomo, J. Gutiérrez-Angonese, C. ArnaizSchmitz, C. Montes, and F. Narváez. 2019. Assessing the effects of past and future land cover changes in ecosystem services, disservices and biodiversity: A case study in Barranquilla Metropolitan Area (BMA), Colombia. Ecosystem Services 37: 100915.

Wetland Science & Practice October 2020 291


Aponte, H. and A. Cano. 2013. Estudio florístico comparativo de seis humedales de la costa de Lima (Perú): actualización y nuevos retos para su conservación. Revista Latinoamericana de Conservación 3(2): 15-17. Bishop, M., S. Powers, H. Porter, and C. Peterson. 2006. Benthic biological effects of seasonal hypoxia in a eutrophic estuary predate rapid coastal development. Estuarine Coastal Shelf Science 70(3): 415-422. Boyer, T. and S. Polasky. 2004. Valuing urban wetlands: A review of non-market valuation studies. Wetlands 24(4): 744–755. Campos-Caba, R. 2016. Análisis de marejadas históricas y recientes en las costas de Chile. Memoria del proyecto para optar al Título de Ingeniero Civil Oceánico, Facultad de Ingeniería, Universidad de Valparaíso, Valparaíso, Chile. 136. Carvajal, M., M. Contreras-Lopez, P. Winckler, and I. Sepúlveda. 2017. Meteotsunamis occurring along the southwest coast of South America during an intense storm. Pure and Applied Geophysics 174(8): 3313–3323. Cepal Comisión Económica para América Latina y el Caribe. 2019. Informe de avance cuatrienal sobre el progreso y los desafíos regionales de la Agenda 2030 para el Desarrollo Sostenible en América Latina y el Caribe. Online https://www.cepal.org/es/ publicaciones/44551-informe-avance-cuatrienal-progreso-desafiosregionales-la-agenda-2030-desarrollo Darrah, S., Y. Shennan – Farpón, J. Loh, N. Davidson, C. Finlayson, C. Royal, M. Gardner, and M. Walpole. 2019. Improvements to the Wetland Extent Trends (WET) index as a tool for monitoring natural and human-made wetlands. Ecological Indicators 99: 294-298. Davidson, N.C. 2014. How much wetland has the world lost? Longterm and recent trends in global 295 wetland area. Marine and Freshwater Research 65 (10): 934–941. Davidson, N.C., E. Fluet-Chouinard, and C.M. Finlayson. 2018. Global extent and distribution of wetlands: trends and issues. Marine and Freshwater Research 69: 620–627. Etter, A., A. Andrade, K. Saavedra, P. Amaya, and P. Arévalo. 2017. Estado de los ecosistemas colombianos: Una aplicación de la metodología de la Lista Roja de Ecosistemas ver. 2.0. Informe Final. Bogotá, Colombia: Pontificia Universidad Javeriana. Faulkner, S. 2004. Urbanization impacts on the structure and function of forested wetlands. Urban Ecosystem 7(2): 89–106. Gallant, A. 2015. The Challenges of Remote Monitoring of Wetlands. Remote Sensing 7(8): 10938-10950. González, S., K. Yáñez-Nevea, and M. Muñoz. 2014. Effect of coastal urbanization on Sandy beach coleóptera Phaleria maculata (Kulzer, 1959) in northern Chile. Universidad Católica del Norte. Coquimbo, Chile. Marine Pollution Bulletin 83 (1): 265-274. Hallegatte, S., C. Green, R.J. Nicholls, and J.Corfee-Morlot. 2013. Future flood losses in major coastal cities. Nature Climate Change 3: 802–806. Hidalgo, R., F. Arenas, and D. Santana. 2016. ¿Utópolis O Distópolis? Producción Inmobiliaria y Metropolización en el litoral central de Chile (1992-2012). Revista EURE 42 (126): 27-54. Jaramillo, U., J. Cortés-Duque, and C. Flórez. 2015. Colombia Anfibia, un país de humedales. Volumen 1, Instituto Alexander von Humboldt. Instituto Alexander von Humboldt, Bogotá, Colombia. Online https://doi.org/10.1007/s13398-014-0173-7.2

292 Wetland Science & Practice October 2020

Hettiarachchi, M., T. H. Morrison and C. McAlpine. 2015. Fortythree years of Ramsar and urban wetlands. Global Environmental Change 32: 57-66. Gibril, M. B., B. Kalantar, R, Al-Ruzouq, N. Ueda, V. Saeidi, A. Shanableh, and H.Z. Shafri. 2020. Mapping heterogeneous urban landscapes from the fusion of digital surface model and unmanned aerial vehicle-based images using adaptive multiscale image segmentation and classification. Remote Sensing 12(7):1081. Kim, Y., D.A. Eisenberg, E.N. Bondank, M.V. Chester, M, G. Mascaro, and B.S. Underwood. 2017. Fail-safe and safe-to-fail adaptation: decision-making for urban flooding under climate change. Climatic Change 145(3–4): 397–412. Maltby E. and M. Acreman. 2011. Ecosystem services of wetlands: pathfinder for a new paradigm. Hydrological Sciences Journal 56:8, 1341-1359. McInnes, R. 2014. Recognising wetland ecosystem services within urban case studies. Marine and Freshwater Research 65: 575–588. Mitsch, W. and J. Gosselink. 2015. Wetlands. Fifth Edition. John Wiley & Sons, New Jersey, USA. Narayan, S., M.W. Beck, P. Wilson, C.J. Thomas, A. Guerrero, C.C. Shepard, B.G. Reguero, G. Franco, J.C. Ingram, and D. Trespalacios. 2017. The value of coastal wetlands for flood damage reduction in the northeastern USA. Scientific Reports 7(1): 9463 - 9462. Martínez, C. P. López, C. Rojas, J. Quense, R. Hidalgo, and F. Arenas. 2020. A sustainability index for anthropized and urbanized coasts: the case of Concón Bay, central Chile. Applied Geography 116. Martínez, C., M. Quezada, and P. Rubio. 2011. Historical changes in the shoreline and littoral processes in a headland bay beach in central Chile. Geomorphology 135: 80-96. Martínez, C., M. Contreras-López, P. Winckler, H. Hidalgo. E. Godoy, and R. Agredano. 2018. Coastal erosion in central Chile: a new hazard? Ocean and Coastal Management 156: 141-155. MINAM - Ministerio del Ambiente, Perú. 2015. Estrategia Nacional de Humedales. Aprobada por Decreto Supremo Nº 004-2015-MINAM. MINAM - Ministerio del Ambiente, Perú. 2019. Mapa Nacional de Ecosistemas del Perú. Online https://sinia.minam.gob.pe/mapas/ mapa-nacional-ecosistemas-peru Moschella, P. 2018. Peri-urbanization and land management sustainability in Peruvian cities. Geography. Université de Strasbourg. Online https://tel.archives-ouvertes.fr/tel-02144701 Moschella, P. 2012. Variación y protección de humedales costeros frente a procesos de urbanización: casos Ventanilla y Puerto Viejo. (Tesis) Pontificia Universidad Católica del Perú, Lima. Patino, J.E., and L.M. Estupinan-Suarez. 2016. Hotspots of wetland area loss in Colombia. Wetlands 36: 935–943. Pintos, P., and A. Sgroi. 2012. Efectos del urbanismo privado en humedales de la Cuenca baja del río Luján, provincial de Buenos Aires, Argentina. Estudio de la megaurbanización San Sebastián. Augmdomus 4: 25-48. Pulido, V. M., and L. Bermúdez. 2018. Estado actual de la conservación de los hábitats de los Pantanos de Villa, Lima, Perú. Arnaldoa 25(2): 679-702. Rangel-Buitrago, N., M. Contreras-López, C. Martínez, and A. William. 2018. Can coastal scenery be managed? The Valparaíso region, Chile as a case study. Ocean and Coastal Management 163: 383-400.


Ramírez, D. W., and H. Aponte. 2018. Por qué los Humedales de Puerto Viejo perdieron su protección legal: analizando los motivos. Revista peruana de biología 25(1): 49-54. Ramsar 2018. Urban wetlands: prized land not wasteland. Online http://www.worldwetlandsday.org Ramsar. 2019. Humedales: Una solución natural al cambio climático. Online http://www.worldwetlandsday.org Ricaurte, L.F., M.H. Olaya-Rodríguez, J. Cepeda-Valencia, D. Lara, J. Arroyave-Suárez, C. Max Finlayson, and I. Palomo. 2017. Future impacts of drivers of change on wetland ecosystem services in Colombia. Global Environmental Change 44: 158–169. Ricaurte, L.F., and others. 2019. A classification system for Colombian wetlands: an essential step forward in open environmental policy-making. Wetlands 39: 971-990. Rivillas-Ospina, G.D., G. Ruiz-Martínez, R. Silva, E. Mendoza, C. Pacheco, G. Acuña, J. Rueda, A. Felix, J. Pérez, and C. Pinilla. 2018. Physical and morphological changes to wetlands induced by coastal structures. In: C.W. Finkl and C. Makowski (eds.). Coastal Wetlands: Alteration and Remediation. Springer, Cham, Switzerland. pp. 275315. Rodríguez, M. (ed.). 2015. ¿Para dónde va el Río Magdalena? Riesgos sociales, ambientales y económicos del proyecto de navegabilidad. Foro Nacional Ambiental, Bogotá, Colombia. Rodríguez, M. 2017. Variación de humedales costeros e irrigaciones agrícolas: el caso de la Albúfera de Medio Mundo y el área agrícola de Huaura. (Tesis) Pontificia Universidad Católica del Perú, Lima. Rojas, C., J. Pino, C. Basnou, and M. Vivanco. 2013. Assessing land use and cover changes in relation to geographic factors and urban planning in the Metropolitan Area of Concepción (Chile). Applied Geography 39: 93–103. Rojas, C. 2018. Desafíos en la Planificación Territorial: Humedales Urbanos una oportunidad de gestión y participación para ciudades más sustentables y resilientes. En: Ministerio de Medio Ambiente. La Vía Medio Ambiental. Desafíos y Proyecciones para un Chile Futuro, 191-201.

Rojas, C., J. Munizaga, O. Rojas, C. Martínez, and J. Pino. 2019. Urban development versus wetland loss in a coastal Latin American city: lessons for sustainable land use planning. Land Use Policy 80: 47 - 56. Rojas, O., M. Mardones., J. Arumí, and M. Aguayo. 2014. Una revisión de inundaciones fluviales en Chile, período 1574–2012: Causas, recurrencia y efectos geográficos. Revista Geográfica Norte Grande 57: 177–192. Patino, J. and L. Estupinan-Suarez. 2016. Hotspots of wetland area loss in Colombia. Wetlands 36: 935–943. Seto, K. C., M. Fragkias, B. Güneralp, and M.K. Reilly. 2011. A meta-analysis of global urban land expansion. PLoS One 6(8): e23777. Sousa, P.T. Jr., M.T, Fernandez Piedade, and E. Candotti. 2011. Brazil’s forest code puts wetlands at risk. Letter to Nature. Nature 478: 458. United Nations. 2011. World Urbanization Prospects: The 2011 Revision, Highlights (ST/ESA/SER.A/322). New York, United States. http://www.un.org/en/development/desa/population/publications/pdf/ urbanization/WUP2011_Report.pdf. United Nations. 2014. World Urbanization Prospects: The 2014 Revision, Highlights (ST/ESA/SER.A/352). New York, United States. http://doi.org/10.4054/DemRes.2005.12.9. United Nations. 2018. Revision of World Urbanization Prospects. Online https://www.un.org/development/desa/publications/2018revision-of-world-urbanization-prospects.html. UN HABITAT. 2017. New Urban Agenda. Quito. Online http://habitat3.org/wp-content/uploads/NUA-English.pdf. Winckler, P., M. Contreras, J. Beyá, and R. Campos-Caba., 2017. El temporal del 8 de agosto de 2015 en la región de Valparaíso, Chile Central. Latin American Journal of Aquatic Research 45(4): 622-648. Wittmann, F., E. Householder, A. Lopes, A. de Oliveira, J. Wolfgang, and M. Piedade. 2015. Implementation of the Ramsar Convention on South American wetlands: an update. Research and Reports in Biodiversity Studies 4: 47 – 58.

Wetland Science & Practice October 2020 293


WETLAND CONSERVATION/EDUCATION/OUTREACH

Wetland Conservation Concerns in Southern Mexico Tatiana Lobato de-Magalhães, Ph.D., PWS, Lucia Guadalupe Cruz, and Everardo Barba1

ABSTRACT pproximately 16 percent of Southern Mexico’s surface area is comprised of wetlands which harbor an abundance of plant and animal species, including endangered and endemic species. With two-thirds of the total wetlands of Mexico and one-third of Mexican Ramsar sites, the Southern Mexico region plays a critical role in wetland conservation worldwide. Despite national and international efforts, many wetland species and ecosystems are threatened in this region. This review includes information related to seven Southern Mexico states: Campeche, Chiapas, Guerrero, Oaxaca, Quintana Roo, Tabasco, and Yucatán. From coastal areas to highlands, this region has around 2,020 mapped wetlands (64,298 km2) and 41 Ramsar sites. Alarmingly, only 13 of the 41 Ramsar sites have management plans implemented. Regardless of the importance of inland wetlands in terms of their area and economic value, issues regarding their conservation and restoration are generally lacking or neglected. Southern Mexican wetlands are also severely threatened by changes in natural habitats, particularly those associated with excessive exploitation of natural resources, tourism, and the oil industry.

A

INTRODUCTION Southern Mexico is a megadiverse, neotropical region that harbors several types of wetlands such as mangroves, riparian forests, floodplains, and cenotes (sinkholes) (Figure 1). According to the National Wetland Inventory, wetlands cover six percent of Mexico (CONAGUA 2020). In Southern Mexico, 2,020 wetlands occupy 64,298 km2, representing two-thirds of the total wetlands in Mexico and 16% of the Southern Mexican States territory (the states of Campeche, Chiapas, Guerrero, Oaxaca, Quintana Roo, Tabasco, and Yucatán). Mexico is the country with the second highest number of Ramsar sites (142 sites designated as Wetlands of International Importance), behind the United Kingdom (175 sites) (Mauerhofer et al. 2015). There are 41 Ramsar sites in Southern Mexico, which represents approximately 29% of the total Mexican Ramsar sites (Ramsar 2020a; 1 El Colegio de La Frontera Sur (ECOSUR), Mexico; Corresponding author contact: ebarba@ecosur.mx

294 Wetland Science & Practice October 2020

Table 1). The region’s wetlands are valuable heritage places that provide several ecosystem services (Smardon 2006; Gortari-Ludlow et al. 2015), and substantially contribute to maintaining biodiversity at local and landscape levels (Mora-Olivo et al. 2013; Alcocer and Aguilar-Sierra 2019). This region has some of the highest levels of aquatic plant species richness and endemism worldwide (Murphy et al. 2019). The National Wetland Policy highlights Tabasco State as having a vast expanse of wetlands (floodplain zones), particularly the Pantanos de Centla Biosphere Reserve (Figure 1a, Figure 2), with an area of 3,027 km2, covering 12% of the total state surface. Among these, the Grijalva River and the Usumacinta River form an estuarine region which is considered one of the most important deltas in North and Mesoamerica because of the water flow and the importance for migratory birds and other species (IUCN 2020; SEMARNAT 2020). WETLAND TYPES AND CLASSIFICATION According to the National Wetland Inventory (CONAGUA 2020), Southern Mexican wetlands are grouped into three major classes: 1) marine and coastal wetlands, including marine and estuarine systems, 2) inland wetlands, including lacustrine, palustrine, and riverine systems, and 3) humanmade wetlands (Table 1; Figure 3). Wetlands are further classified based on their hydrological regime (permanent, intermittent, and temporary wetlands), soil properties (texture and composition), and vegetation type, such as the endemic flooded low evergreen forest ecosystem in the Yucatán peninsula (Bala’an K’aax) and the islands of vigorous tree vegetation associated with springs and water holes, which constitute a critical habitat for wildlife, Los Petenes in the Campeche State (Figure 1b) (Lot 2004; Ramsar 2020a). Marine and Coastal Wetlands This wetland type represents around 15% of the total mapped wetlands. Marine wetlands (Figure 1c) are most represented by a seagrasses community or ceibadal (e.g., Halodule, Syringodium, and Thalassia species) (Creed et al. 2003) and mangroves (Figure 1d). Some threatened mangrove plant species are Avicennia germinans (black


mangrove), Conocarpus erectus (button mangrove), Laguncularia racemosa (white mangrove), and Rhizophora mangle (red mangrove) (SEMARNAT 2010). Campeche and Quintana Roo States have a large portion of mangroves on the Atlantic Coast, and Chiapas State contains large areas of mangrove on the Pacific Coast. Inland Wetlands Most of the Southern Mexico wetlands are classified as inland wetlands (82%). They include mostly freshwater palustrine wetlands described as swamps, floodplains, marshes, and forested wetlands (riparian forests, palm thickets, and inundated low lands) (Figure 1e). Riparian forests are comprised of Salix negra, S. caroliniana, and S. chile (willows). Lowlands floodable forests are represented by Annona glabra (swamp apple) (Campeche State), Dalbergia brownei (rosewood), and Ficus padofolia (fig tree) (Tabasco State) (Lot 2004). Lacustrine wetlands (Figure 1f) occur mostly in highlands and are less abundant than palustrine wetlands (Olmsted 1993). Rooted floating-leaved plants are numerous in lakes, lagoons, canals, and open freshwater wetlands (e.g., Nymphaea – waterlily, Nuphar – waterlily, Nymphoides – floatingheart, Potamogeton – pondweed, and Sagittaria – arrowhead) (Lot 2004). Fifty-eight plant species are associated with calcareous warm-water rivers of Yucatán peninsula including Bacopa monnieri (water hyssop), Eleocharis geniculata (spikesedge), Hydrocotyle umbellata (manyflower), Lemna aequinoctialis (duckweed), Nymphaea ampla (waterlily), Paspalum notatum (bahiagrass), and Typha domingensis (cattail) (Tapia-Grimaldo et al. 2017). Sinkholes (cenotes) are a unique type of inland wetland associated with a karstic geology. The cenotes are an important freshwater resource in the Yucatán peninsula region that are highly impacted by tourism (Figure 1g); they harbor several rare and threatened aquatic species (Cervantes-Martínez

FIGURE 1. Southern Mexico wetlands: (a) palustrine wetland, El Palmar, Pantanos de Centla Biosphere Reserve, Tabasco State; (b) islands of tree vegetation associated with springs and water holes, Los Petenes, Campeche State; (c) marine wetland with seagrasses vegetation, Laguna de Términos, Campeche State; (d) mangrove, Tabasco State; (e) riverine rainforest, Rio Tzendales, Chiapas State; (f) lacustrine wetland, Laguna Catazaja, Chiapas, (g) sinkhole (cenote) frequented by tourists, Quintana Roo State, and (h) wetland habitats associated with waterfalls, Parque Nacional Cañón del Sumidero, Chiapas State.

(Photos a, b, c, e, f courtesy of Everardo Barba; d courtesy of Alejandro Betancourth; g and h courtesy of Paula Montoya)

Wetland Science & Practice October 2020 295


FIGURE 2. Panoramic view of Pantanos de Centla Biosphere Reserve, Tabasco State.

(Photo courtesy of the Mexico Mangrove Monitoring System developed by CONABIO/ SEMARNAT. Photo taken by Joanna Acosta)

FIGURE 3. Wetlands in Southern Mexico States. (Adapted from National Wetland Inventory; CONAGUA 2020)

(Map elaborate by Tatiana Lobato de Magalhães Ph.D., PWS)

296 Wetland Science & Practice October 2020

et al. 2018; Mondragón-Mejía et al. 2019). Highland wetlands are extremely important for the provision of water to Southern Mexico cities like San Cristóbal de las Casas. The dominant aquatic plants in highland wetlands are Typha (cattail), Phragmites (common reed or carrizal), Cyperus (umbrella sedges), Eleocharis (spike-rushes), and Schoenoplectus (bulrushes) (Lot 2004; Chediack et al. 2018). Human-made Wetlands Human actions have created wetlands in places, especially through dam construction. They represent about three percent of Southern Mexico wetlands. ECOSYSTEM FUNCTIONS AND THREATS Southern Mexican wetlands provide several ecosystem services and are also impacted by human uses such as livestock, aquaculture, excessive exploitation of natural resources, and industrial expansion (Tables 2 and 3). For example, Ría Lagartos (Yucatán State) is an important estuarine wetland for flamingo nesting as well for economic activities such as fishing, agriculture, salt production, and livestock. Another activity that has a strong impact on wetlands is the selective extraction of native palms such as Pseudophoenix sargentii (buccaneer palm or kuka’), Thrinax radiata (thatch palm or chit), and Coccothrinax readii (Mexican silver palm or nacax), which are used for decoration along avenues and hotels in cities like Cancun. In Laguna de Terminos (Figure 1b) (Campeche State), the exploitation of natural resources has been crucial for the local economy during the last three centuries, through the extraction of dye sticks, precious woods, and chewing gum. This wetland is known for the sustainable use and management of Crocodylus moreletti (Mexican crocodile) populations for commercial purposes based on its skin (SEMARNAT 2020).


Regarding economic valuation of wetland ecosystem services, inland wetlands are rated higher than estuarine ones in Tabasco State: palustrine ($9,689 USD/ha/year), lacustrine ($6,366), mangrove ($2,653), and coastal lagoon ($1,926) (Camacho-Valdez et al. 2020). Overall, inland wetlands are threatened by the land-use changes and by the oil industry, especially the swamps, floodplains, and marshes in Tabasco State. The extraction of hydrocarbon has led to major wetland impacts (Domínguez-Domínguez et al. 2019; Camacho-Valdez et al. 2020). Furthermore, in the last two decades the Pantanos de Centla (Tabasco State) has experienced a notable land-use change – the conversion of natural floodplain vegetation to livestock and agricultural areas (De la Rosa-Velázquez et al. 2017). Lowland floodable forests have been drastically reduced by agricultural activity (conversion to pasture and farmland) and overexploitation of the Haematoxylum campechianum (campeachy tree or logwood), which was used for a long time as a natural source of textiles dye, applied in histology for staining, and for medicinal uses (Lot 2004). Frequent threats to the highland wetlands are urbanization, pollution, mining, and agricultural activities. Several monocotyledons aquatic species historically recorded above 2,000 m a.s.l. (Chiapas State) were not detected in a recent floristic study; the authors consider that it could indicate a process of local extinction (Chediack et al. 2018).

BIODIVERSITY, ENDEMISM, AND THREATENED SPECIES Southern Mexico wetlands harbor several endemic species such as Lithobates brownorum (leopard frog), Bolitoglossa yucatana (Yucatan mushroomtongue salamander), Cyprinodon macularius (desert pupfish), Caretta caretta (loggerhead sea turtle), Chelonia mydas (green sea turtle), Eretmochelys imbricada (hawksbill sea turtle), and Sanopus splendidus (splendid toadfish). The following states harbor a rich number of aquatic plant species and are considered priority states for the conservation of strictly aquatic plant species in Mexico: Chiapas (225 species), Campeche (220), Oaxaca (210), and Tabasco (186) (Mora-Olivo et al. 2013). The Pantanos de Centla (Tabasco State) alone harbors around 569 plant species (76 used by people and 13 rare or threatened) and a fauna with more than 523 vertebrate species (IUCN 2020; SEMARNAT 2020). The Anillo de Cenotes are home to endemic species of reptiles (e.g., Terrapene carolina yucatana – Yucatan box turtle), amphibians (e.g., Bolitoglossa yucatana), and birds (e.g., Stelgidopteryx ridgwayi – Yucatan rough-winged swallow, Cyanocorax yucatanicus – Yucatan jay, and Melanoptila glabirostris – black catbird). These cenotes are also home to a number of endangered or threatened species (Cervantes-Martínez et al. 2018; IUCN 2020; Ramsar 2020a). The highland wetland Humedales de Montaña La Kisst (Chiapas, 2,120 m a.s.l.) supports great populations of fish and amphibians, with at

TABLE 1. Extent of wetlands in Southern Mexico based on National Wetland Inventory (NWI) and Ramsar sites.

Wetland Total NWI Marine and Coastal Inland Artificial Ramsar sites Marine and Coastal Inland

Number of Wetlands

Total Area (km2)

Reference

2,020 392 1,558 70 41 26 15

64,298 9,602 52,610 2,085 34,232 29,451 4,782

CONAGUA 2020

Ramsar 2020a

TABLE 2. Ecosystem services provided by wetlands in Southern Mexico. (Adapted from Smardon 2006; Camacho-Valdez et al. 2020; Ramsar 2020a.)

Ecosystem Services Provided by Wetlands • Recreation and tourism • Scientific and educational uses • Heritage places • Drinkable water storage • Hydrological flow regulation • Biological production (wetland food and non-food products) • Biogeochemical cycle regulation (erosion protection, pollution control and detoxification, nutrient cycling, and soil formation) • Wildlife habitat and biodiversity conservation (genetics, endemism, and rare and threatened species) Wetland Science & Practice October 2020 297


FIGURE 4. Ramsar sites in Southern Mexico States. (Adapted from Ramsar 2020a)

(Map elaborate by Tatiana Lobato de Magalhães Ph.D., PWS.)

least 10 species being endemic or under a protection category (e.g., the endemic fish Profundulus hildebrandi - Chiapas killifish, and the endemic plant Wolffia columbiana – Columbian water-meal) (Chediack et al. 2018; Ramsar 2020a). Several wetland species are critically endangered due to habitat loss. The International Union for Conservation of Nature’s Red List of Threatened Species lists one species of Fungi, 46 plants, and 375 animal species associated with Southern Mexican coastal and inland wetlands (IUCN 2020). Concerning the threatened categories of IUCN, seven species are critically endangered, 17 species endangered, 27 species vulnerable, 15 species near threatened, 276 species least concern, and 80 species data deficient (Table 4). In regard to aquatic animal species, there are four species of mollusks, 125 species of arthropods (Insecta and Malacostraca), 145

TABLE 3. Activities that threaten Southern Mexico wetlands. (Adapted from Gortari-Ludlow et al. 2015; Domínguez-Domínguez et al. 2019; CamachoValdez et al. 2020; Ramsar 2020a.)

Wetland Threats • Change in natural habitats (agriculture, livestock, aquaculture, and human settlements) • Excessive exploitation of natural resources (fishing and harvesting aquatic resources, logging and wood harvesting, hunting and collecting terrestrial animals, marine and freshwater aquaculture, and gathering plants) • Changes in flow regime (drainage and canals construction) • Wastewater (rural, urban, and industrial) • Drought (high temperature and high evaporation) • Infrastructure projects (road construction) • Oil industry (hydrocarbon extraction and processing) • Unsustainable use (tourism and navigation) TABLE 4. Threatened species in Southern Mexico wetlands. (Adapted from IUCN 2020.)

Taxonomic Group Fungi Plants Mollusks Arthropoda Fishes Amphibians Reptiles Aves Mammals Total

Critically Endangered 1 1 2 2 1 7

Endangered

Vulnerable

2 7 7 1 17

8 12 5 1 1 27

298 Wetland Science & Practice October 2020

Near Least Concern Threatened 1 43 3 1 72 6 76 3 1 22 6 57 15 276

Data Deficient 1 36 42 1 80

Total 1 46 4 125 145 5 30 64 2 422


fishes, five amphibians, 30 FIGURE 5. Number of Ramsar sites that satisfy each Ramsar criterion for Southern Mexico. Wetland type reptiles, 64 species of wa(criterion 1), biological diversity (criteria 2, 3, and 4), waterbirds (criteria 5 and 6), fishes (criteria 7 and 8), and other taxa (criterion 9). (Adapted from Ramsar 2020a, b) terfowl birds, two mammals Rheomys mexicanus (Mexican water mouse) and Trichechus manatus (American manatee). Among the arthropods there are 152 species of insects that depend on the aquatic systems for critical stages in their life cycles, including dragonflies and damselflies. Coastal ecosystems and mangroves are crucial for the threatened crocodiles and caimans (i.e., Crocodylus moreletti - Mexican crocodile and Caiman crocodilus - spectacled caiman). Southern Mexico also shelters a vast number of waterfowl, and migratory aquatic birds that come to its wetlands in the winter, and several of these species are threatened (Platt et al. 2010; Domínguez-Domínguez et al. 2019; IUCN 2020). (Map elaborate by Tatiana Lobato de Magalhães Ph.D., PWS) The aquatic turtle Dermatemys mawii (white turtle) is the only The first Mexican Ramsar site was designated in critically endangered reptile in the region. Other threatened freshwater aquatic turtles are Kinosternon creaseri (creaser’s 1986 and is located in Southern Mexico at Yucatán State, Reserva de la Biosfera Ría Lagartos. The most recent, mud turtle), K. integrum (Mexican mud turtle), K. oaxacae Humedales de Montaña María Eugenia, was added in 2012 (Oaxaca mud turtle), Trachemys ornata (ornate slider), and and is located in Chiapas state (Ramsar, 2020a). A total T. scripta (pond slider). of six sites were also designated as UNESCO Biosphere RAMSAR SITES Reserves. With regard to geographical distribution, all Southern Mexico has 41 Ramsar sites covering 34,232 seven Southern Mexican states have at least one Ramsar km2 (2% of the total country surface). It represents almost site (Ramsar 2020a) (Figure 4). Quintana Roo state has a third of the 142 total Mexican Ramsar sites. Around the largest number of sites (13), followed by Chiapas (12), 85% (35 wetlands) of Southern Mexico sites occur in low Yucatán (7), Oaxaca (4), and Campeche (3), while Guerrero elevations (< 300 m a.s.l.), while the highest elevation and Tabasco have only one Ramsar site each. Additionally, Ramsar sites are attributed to Humedales de Montaña La two Southern Mexico Ramsar sites wetlands extend into Kisst and Humedales de Montaña María Eugenia, both in the territory of other countries. Parque Nacional Lagunas Chiapas State (2,120 m a.s.l.). The majority of Ramsar sites de Montebello lies on the Southern border, extending into are classified as coastal and marine (64%), followed by Guatemala and Parque Nacional Arrecifes de Xcalak into inland wetlands (36%) (Table 1). Among inland wetlands, Belize border. 13 sites have a permanent water regime and two sites are The Ramsar Convention considers nine criteria to considered seasonal or intermittent. Mexican Ramsar sites designate wetlands as of international interest for conservahave a mean size of 835 km2. The largest site (7,050 km2) is tion (Ramsar 2020a): wetland type (criterion 1), biological Área de Protección de Flora y Fauna Laguna de Términos diversity (criteria 2, 3, and 4), waterbirds (criteria 5 and (Campeche State), while the smallest site (0.2 km2) is Playa 6), fishes (criteria 7 and 8), and other taxa (criterion 9) Barra de la Cruz (Oaxaca State) (Ramsar 2020a). (Ramsar 2020b). Criteria based on wetland type and bioWetland Science & Practice October 2020 299


diversity are more frequently reported on than ones related to specific taxa (Figure 5). Among the 41 Southern Mexico Ramsar sites none satisfy all criteria and one fills only one criterion. The latter site is the Parque Nacional Cañón del Sumidero (criterion 1) that contains a unique example of a natural wetland type (humid habitats associated with waterways and waterfalls) (Figure 1h) and harbors threatened species such as Crax rubra (great curassow), Ateles geoffroyi (black-handed spider monkey), Crocodylus acutus (American crocodile), Leopardus wiedii (margay), and Rinodina chrysomelaena (bright yellow crustose lichen) (IUCN 2020). CONSERVATION PERSPECTIVE AND CONCLUSIONS Several actions have been launched since 1986 when Mexico became a signatory country of the Ramsar Convention, particularly the creation of the National Wetland Policy and a National Wetlands Committee, the designation of 142 Ramsar sites, and development of the National Wetland Inventory (CONAGUA 2020; SEMARNAT 2020). Additionally, since 1936, Mexico has been party to an agreement with the United States for the protection of migratory birds, which has contributed to the implementation of bi-national initiatives that have improved wetland conservation in Mexico. Despite the international importance of the region’s Ramsar wetlands, only 71% of the Southern Mexico Ramsar sites have management plans (25 sites with concluded management plans, four sites with plans in preparation). Of these sites, however, only 13 have plans that have already been implemented. Surprisingly, the latter number represents 50% of the total Ramsar implemented management plans in Mexico. These findings denote a low level of concern with practical actions to conserve wetlands in the country (Ramsar 2020a). Further, even the implementation of management plans does not necessarily promote concrete conservation actions to protect wetlands (GortariLudlow et al. 2015). Overall, inland wetlands are being overlooked and not getting the attention they need for conservation and restoration, not only in Mexico but around the world (Reis et al. 2017). Environmental policies in Mexico have been more focused on mangroves and coastal wetlands than on inland wetlands. In 2003, the government approved a federal law regulating the wise-use, conservation, and restoration of coastal wetlands and mangroves (Norma Oficial Mexicana 022/2003; SEMARNAT 2003), and in 2005, established a National Mangrove Committee (CONABIO 2020). Unsurprisingly, estimates suggest that inland wetland losses are larger and faster than losses in coastal wetlands (Davidson 2014). Despite freshwater ecosystems recording around

300 Wetland Science & Practice October 2020

90% of the aquatic plant species richness of Mexico (MoraOlivo et al. 2013) and having higher economic values than estuarine wetlands (Camacho-Valdez et al. 2020), Mexican conservation actions focused on inland wetlands are lacking, particularly for highland wetlands (Alcocer and Aguilar-Sierra 2019). Strong synergies among stakeholders that engage the population, private and governmental sectors, decisionmakers, non-governmental organizations, and the academy are crucial to improving wetland conservation in Mexico. Projects that integrate science and practice are also essential for wetland restoration and conservation at local and regional levels. Through improved efforts to increase wetland protection awareness and acquisition of more detailed data on the degradation and change of wetland areas, including risk assessment analysis (Camacho-Valdez et al. 2020), Southern Mexican wetlands could achieve a positive future scenario. n ACKNOWLEDGMENTS We thank K. Murphy, D. Jones, and D. Orsini for reviewing this manuscript. The first author holds a postdoctoral fellowship from ECOSUR (El Colegio de la Frontera Sur) at CONACyT (National Council of Science and Technology) in Mexico. We appreciate the valuable comments of the editor R. Tiner. REFERENCES

Alcocer, J. and V. Aguilar-Sierra. 2019. Biodiversity in inland waters. Chapter 2. In: A.L. Ibáñez (ed.) Mexican Aquatic Environments: A General View from Hydrobiology to Fisheries. Springer Nature Switzerland. pp 43-75. Doi:10.1007/978-3-030-11126-7 Camacho-Valdez, V., A. Saenz-Arroyo, A. Ghermandi, D.A. NavarreteGutiérrez, and R. Rodiles-Hernández. 2020. Spatial analysis, local people’s perception and economic valuation of wetland ecosystem services in the Usumacinta floodplain, Southern Mexico. PeerJ 8: e8395. Doi:10.7717/peerj.8395. Cervantes-Martínez, A., M.A. Gutiérrez-Aguirre, M. Elías-Gutiérrez, A.M. Arce-Ibarra, and A. García-Morales. 2018. Aquatic biodiversity in cenotes from the Yucatán Peninsula (Quintana Roo, Mexico). Teoría y Praxis 14(25): 49-68. Chediack, S.E., N. Ramírez-Marcial, M. Martínez-Icó, and H.E. Castañeda-Ocaña. 2018. Macrófitos de los humedales de montaña de San Cristóbal de Las Casas, Chiapas, México. Revista mexicana de biodiversidad 89(3): 757-768. Doi:10.22201/ib.20078706e.2018.3.2420 CONABIO. 2020. Comité Nacional de Manglares. Available via https:// www.biodiversidad.gob.mx/monitoreo/smmm/comiteNacional Accessed 17 Mar 2020. CONAGUA. 2020. Visualizador de humedales de la República Mexicana - Inventario Nacional de Humedales. Available via https://www.gob.mx/ conagua/acciones-y-programas/visualizador-de-humedales-de-la-republica-mexicana-inventario-nacional-de-humedales Accessed 15 Mar 2020. Creed, J.C., R.C. Phillips, and B.I. Van Tussenbroek. 2003. The seagrasses of the Caribbean. In: E. P. Green and F.T. Short (Eds.) World Atlas of Seagrasses. Cambridge: UNEP-WCMC. pp. 234-242.


Davidson, N.C. 2014. How much wetland has the world lost? Longterm and recent trends in global wetland area. Marine and Freshwater Research 65(10): 934-941. Doi: 10.1071/MF14173

Olmsted, I. 1993. Wetlands of Mexico. In: Whigham, D.F., Dykyjová, D., and Hejný, S. (Eds.) Wetlands of the World I: inventory, ecology and management, Sprinter. pp. 637-677. Doi:10.1007/978-94-015-8212-4.

De la Rosa-Velázquez, M.I., A. Espinoza-Tenorio, M.Á. Díaz-Perera, A. Ortega-Argueta, R. Ramos-Reyes, and I. Espejel. 2017. Development stressors are stronger than protected area management: a case of the Pantanos de Centla Biosphere Reserve, Mexico. Land Use Policy 67: 340-351. Doi:10.1016/j.landusepol.2017.06.009

Platt, S.G., L. Sigler, and T.R. Rainwater. 2010. Morelet’s Crocodile Crocodylus moreletii. In: Manolis, S.C. and Stevenson, C. (Eds) Crocodiles. Status Survey and Conservation Action Plan. 3.ed, Group Darwin, pp. 79-83.

Domínguez-Domínguez, M., J. Zavala-Cruz, J.A. Rincón-Ramírez, and P. Martínez-Zurimendi. 2019. Management strategies for the conservation, restoration and utilization of mangroves in Southeastern Mexico. Wetlands 39(5): 1-13. Doi:10.1007/s13157-019-01136-z IUCN. 2020. The International Union for Conservation of Nature’s Red List of Threatened. Available https://www.iucnredlist.org/ Accessed 10 Mar 2020. Gortari-Ludlow, N., G. Espinosa-Reyes, J. Flores-Rivas, J. SalgadoOrtiz, and L. Chapa-Vargas. 2015. Threats, conservation actions, and research within 78 Mexican non-coastal protected wetlands. Journal for Nature Conservation 23: 73-79. Lot, A. 2004. Flora and vegetation of freshwater wetlands in the coastal zone of the Gulf of Mexico. In: Caso, M., Pisanty, I., Ezcurra, E., Withers, K., and Nipper, M. (Eds.) Environmental Analysis of the Gulf of Mexico, INECOL, INE, Harte Research Institute, Texas A&M Corpus Christi. Available https://www.harteresearchinstitute.org/environmentalanalysis-gulf-mexico Accessed 04 Feb 2020. Mauerhofer, V., R.E. Kim, and C. Stevens. 2015. When implementation works: A comparison of Ramsar Convention implementation in different continents. Environmental Science & Policy 51: 95-105. Doi:10.1016/j. envsci.2015.03.016 Mondragón-Mejía, J.A., F. Enseñat-Soberanis, and R. Blanco-Gregory. 2019. La percepción de multitud como indicador de gestión sostenible de los cenotes de uso turístico en Yucatán, México. Available http://riull.ull. es/xmlui/handle/915/17850 Accessed 04 Apr 2020 Mora-Olivo, A., J. Villaseñor, and M. Martínez. 2013. Las plantas vasculares acuáticas estrictas y su conservación en México. Acta Botánica Mexicana 103: 27-63. Doi:10.21829/abm103.2013.50

Ramsar. 2020a. Ramsar sites information service. Available via https:// rsis.ramsar.org/ris-search/332?pagetab=0 Accessed 2 Mar 2020. Ramsar. 2020b. The Ramsar Sites Criteria: The nine criteria for identifying Wetlands of International Importance. Available via https://www. ramsar.org/sites/default/files/documents/library/ramsarsites_criteria_eng. pdf Accessed 02 Mar 2020. Reis, V., V. Hermoso, S.K. Hamilton, D. Ward, E. Fluet-Chouinard, B. Lehner, and S. Linke. 2017. A global assessment of inland wetland conservation status. Bioscience 67(6): 523-533. Doi:10.1093/biosci/bix045 SEMARNAT. 2003. Norma Oficial Mexicana 022/2003. Available via http://siga.jalisco.gob.mx/assets/documentos/normatividad/nom022semarnat2003.htm Accessed 17 Mar 2020. SEMARNAT. 2010. Especies nativas de México de flora y fauna silvestres - Categorías de riesgo y especificaciones para su inclusión, exclusión o cambio - Lista de especies en riesgo). NOM-059- ECOL-2010. Diario Oficial de la Federación, México, D. F. (In Spanish) SEMARNAT. 2020. Política Nacional de Humedales. Available via https://agua.org.mx/wp-content/uploads/2017/07/politica-nacional-dehumedales.pdf Accessed 17 Mar 2020. Smardon, R.C. 2006. Heritage values and functions of wetlands in Southern Mexico. Landscape and Urban Planning 74(3-4): 296–312. Doi:10.1016/j.landurbplan.2004.09.009 Tapia-Grimaldo, J., M.T. O’Hare, M.P. Kennedy, T.A. Davidson, J. Bonilla‐Barbosa, B. Santamaría‐Araúz, ... and K.J. Murphy. 2017. Environmental drivers of freshwater macrophyte diversity and community composition in calcareous warm‐water rivers of America and Africa. Freshwater Biology 62(9): 1511-1527. Doi:10.1111/fwb.12962

Murphy, K., A. Efremov, T.A. Davidson, E. Molina-Navarro, K. Fidanza, T.C.C. Betiol, ... and M. Kennedy. 2019. World distribution, diversity and endemism of aquatic macrophytes. Aquatic Botany 158: 103127. Doi:10.1016/j.aquabot.2019.06.006

Wetland Science & Practice October 2020 301


WETLAND CONSERVATION/EDUCATION/OUTREACH

Propagation of Endangered Aquatic Plants: An Experience that Promotes ex situ Conservation and Environmental Education Sandra Nayeli González Mateos1

ABSTRACT etlands are among Mexico’s most threatened habitats as they have not received the degree of protection and conservation as they have in North America. To promote their conservation, the Botanical Garden at the Institute of Biology at the National Autonomous University of Mexico (UNAM), located south of Mexico City, has established a collection of aquatic plants with the intent to have a representative sample of the country’s aquatic plants, with emphasis on the “Cuenca of Mexico”. The collection is being used to propagate aquatic species, conduct research, promote conservation of aquatic plants, and to serve as the foundation for environmental education programs to increase public awareness of these species and the challenges they face.

W

WHY CONSERVE AQUATIC PLANTS IN THE BOTANICAL GARDEN OF THE INSTITUTE OF BIOLOGY – UNAM? In Mexico, aquatic ecosystems are among the country’s most threatened natural habitats. They have received little attention for their conservation and there are few studies related to Mexican aquatic species (Mora-Olivo et al. 2013). Aquatic plants have been linked to man since very ancient times - the element water has been a very important factor for the establishment and development of great civilizations including the great Tenochtitlan in the “Cuenca de México” (the ancient capital of the Aztec empire). Since pre-Hispan-

ic times, aquatic plants have been used for various purposes (Miranda 1980). One of the most important uses was to build the “chinampas” - the agricultural production systems (e.g., “floating gardens”) that helped to shape the city itself and served to “gain” space in the water and form new sections of “land” for the very expansion of the great city (for an example, see https://aztecexplorers.com/2018/07/26/ travelling-in-time-exploring-the-chinampas-of-tlahuacmexico-city/). Species such as Lilaeopsis schaffneriana, Schoenoplectus tabernaemontani, Hydrocotyle ranunculoides, Polygonum amphibium, Lemna gibba and Bidens aurea were used specifically for the construction of “chinampas” (Lot and Novelo 2004). Despite their historic significance, there are currently very few studies for these types of aquatic plants in both natural populations and ex situ conditions. The Aquatic Plant Collection (APC) is a botanical collection of the few of its kind that exist in Mexico and its goal is to have a representative sample of the aquatic vascular plants of the flora of our country, with emphasis on the “Cuenca of Mexico.” WHO ARE WE? The Botanical Garden of the Institute of Biology - UNAM (JB IB-UNAM) is located south of Mexico City and occupies an area of 2.75 hectares (Caballero Nieto 2012). The Aquatic Plant Collection (APC) is distributed in 17 ponds lo-

FIGURE 1. Panoramic view of one of the ponds in the Aquatic Plant Collection containing plants with medicinal and/or food use (e.g., Equisetum hyemale and Nymphaea mexicana). (Photo by Surya Ivonne González Jaramillo.)

1 Botanical Garden of the Institute of Biology –National Autonomous University of Mexico (UNAM), Mexico City, Mexico; author contact: nayelig@ib.unam.mx

302 Wetland Science & Practice October 2020


cated in different sections of the botanical garden and covers an area of approximately 473 m2 (Figure 1). It also contains a greenhouse of 95 m2 for the propagation, growth and development of different species of aquatic plants (Figure 2). Originally, the founders and collaborators of the JB IB-UNAM made various collections for the production of species that show the plant diversity of Mexico. Historical records indicate that in some of the botanical expeditions, aquatic plants were collected (Equisetum hyemale var. affine, Equisetum giganteum, Hydrocotyle sp., Sagittaria sagittifolia and Typha sp. among others) however at that time no specific collection was consolidated strictly for aquatic plants. From 2006, on the initiative of Dr. Javier Caballero Nieto (then Head of the Botanical Garden) all the Garden’s collections were inventoried, strengthened and redefined into two different types of collections according to their composition and function: taxonomic collections and thematic collections. It was under this concept, that the Aquatic Plant Collection (APC) was formalized as a thematic collection. At the invitation of Dr. Javier Caballero, Dr. Antonio Lot Helgueras (specialist researcher in aquatic plants of the Institute of Biology UNAM) and landscape architect Pedro Camarena developed a project for the establishment of this collection. For this project, a theme was defined for each of the existing ponds that included a plant palette of aquatic species and an educational proposal for the use of the collection; this was accomplished by biologist Teodolinda Balcázar Sol (Coordinator of the Educational Area). From then on biologist José Luis López was in charge of the maintenance of the ponds and the asexual propagation counting for it, with the support of the gardener Jesús Rebollo. Some of the collection sites of the plant species were the states of Michoacán, Morelos, Hidalgo and Jalisco. NEW FLIGHT PLAN: ACTIONS TO BOOST APC Starting in 2015, I was appointed curator responsible for the APC to continue developing the master plan designed by Dr. Antonio Lot Helgueras. Each pond shows the different life forms, endemism, introduced species (naturalized species are shown, i.e. those that have dispersed to new environments and have succeeded in these sites and also those species considered weeds, due to the excessive growth of their populations that harms other species), as well as the species used as medicine, food, fibers for the elaboration of handicrafts and building materials. Species used for ornamental (various types and colors of flowers) and ceremonial purposes are also included. After analyzing the conditions of the APC and the greenhouse, it was determined that the main actions to be taken would be: 1) boost the sexual spread of species, 2) conduct different studies to learn more about the development of species

FIGURE 2. Aquatic plant propagation tank in the greenhouse. (Photo by Surya Ivonne González Jaramillo.)

FIGURE 3. Students working on various projects: Surya González (a) working with Nymphaea odorata and university students - Sara Díaz, Isaac Avalos, and Francisco Mendoz (b) working in the greenhouse. (Photos by Diana Ferrusca Domínguez – a, and Nayeli González – b.)

a

b Wetland Science & Practice October 2020 303


under ex situ conditions, and 3) promote environmental education activities for the revaluation of Mexican aquatic plants. These actions have been made possible through multidisciplinary projects that integrate high school students and college undergraduates in research topics related to aquatic plants and even fauna associated with the ponds in the collection. Currently the APC includes 35 species from 19 botanical families with the ultimate goal to increase the number of species and individuals per species. Since 2016, 34 students of high school or university levels have been performing volunteer activities, summer stays, professional internships, social service, or thesis research. Basic horticulture activities are carried out in addition to the design and execution of research that promotes the study of aquatic plant species. This work expands our knowledge about the biology of species, while encouraging students to develop technical skills for hydrophytic plant management and the revaluation of wetlands (Figure 3). SEXUAL REPRODUCTION RESEARCH Studies are carried out with some of the most attractive species of the Family Nymphaceae that are subject to more environmental pressures. The APC includes five species from this family, of which Nymphaea gracilis, N. odorata and N. mexicana (Figure 4) are those for which germination and growth tests have been initiated. These three species were abundant in the “Cuenca of Mexico” and have been elements of the landscape with a great cultural significance (López Martínez 2018). Since their populations have declined or disappeared completely they are now considered as threatened species of extinction, according to current environmental regulations. The JB IB-UNAM contributes

to their ex situ conservation as studies are carried out to understand and improve their propagation. N. gracilis or “atzazamolli” is a water lily with white flowers that stand out from the water level and possesses large green leaves (Figure 4a). Its cultural value is very important as it is represented in the “Florentino Codex” in the section that mentions edible wild plants. This species is endemic to Mexico and, unfortunately, is highly vulnerable to the loss and contamination of water bodies. Nymphaea mexicana called “atlacuetzon” or “paskurinda” inhabited the Mexico Basin and unfortunately its natural populations have declined drastically. This species with bright yellow flowers (Figure 4b) was a very important element in the consolidation of the “chinampas.” It is even represented in pre-Hispanic murals that demonstrate its cultural value. In addition, it had an ornamental use for the Day of the Dead offerings and its leaves were used in the preparation of a special food that was made with the mixture of other elements of the “chinampas” (López Martínez 2018) N. odorata or “apapatla” is a very elegant and showy species whose white flowers float on the surface of the water and offer a very pleasant sweet aroma (Figure 4c). This species was a very important element for the formation of “chinampas” however it has not been seen in the Basin of Mexico since the 1940s (Note: It is the most common water lily in North America). Its leaves are large, which favors the colonization by aquatic fauna, both below water (i.e., for the deposit of snail eggs) and above their leaves (e.g., some small birds perch on them in order to feed on various insects). Since the summer of 2016, individuals of Nymphaea odorata, N. gracilis and N. mexicana have been monitored in the different exhibition ponds and/or in the propagation

FIGURE 4. Three water lilies being studied: Nymphaea gracilis (a), N. mexicana (b), and N. odorata (c). (Photos by APC – a, and Nayeli González – b and c.) a

304 Wetland Science & Practice October 2020

b

c


TABLE 1. Summary of studies on propagation of aquatic plants.

Vegetative propagation Fruits Seeds

N. mexicana

N. odorata

N. gracilis

Successful

Successful

Successful

Not enough have been obtained.

Bee pollination and cross pollination. Successful cross pollination. Abundant fruits in older specimens.

Scarce

Abundant

Abundant

Low germination. Stimulation with constant temperature and dark conditions.

Successful germination. Stimulation with constant temperature conditions.

Germination Viable seeds have not been obtained. Next actions

Fertilization tests for flower and fruit production.

Pre-germination treatments in seeds. Germination in different light and temperature conditions.

FIGURE 5. From left to right: the first two images (with the yellow margin) shows the formation of stoloniferous rhizomes in N. mexicana for vegetative propagation; the central pair of images (with the orange margin) shows the horizontal rhizome of N. odorata that can be cut for production of new individuals and N. odorata observed in fruit that is collected to obtain seeds; the images on the right (with the blue margin) show N. gracilis - its rhizome with small leaves initiating growth and a student performing artificial pollination. (Photos by APC.)

FIGURE 6. Diana Ferrusca separating rhizomes from N. mexicana. (Photo by Surya Ivonne Gonzรกlez Jaramillo.)

FIGURE 7. One of the new species added to the APC collection: Anemopsis californica. (Photo by Nayeli Gonzรกlez.)

Wetland Science & Practice October 2020 305


greenhouse. Sites were periodically reviewed to monitor the formation of flowers and fruits. Fruits were collected and seeds obtained. Species leaflets or rhizomes were also collected to determine favorable characteristics for their spread. In addition, the greenhouse pollination of Nymphaea odorata and N. gracilis produced fruit, viable seeds, and seedlings of both species. Although some seeds of N. mexicana were produced, no seedlings have been established. However N. mexicana did achieve favorable vegetative reproduction by means of rhizomes. There are different factors (e.g., substrate, light, and water level) that influence both germination and the establishment of seedlings so obtaining good-sized individuals from seeds involves a slower process but with greater advantages promoting genetic diversity (Bornette and Puijalon 2011). Table 1 highlights the main results and the next actions with each of the species, while Figure 5 shows the propagation process. PROPAGATING SPECIES UNDER EX SITU CONDITIONS When plants are grown in greenhouses and gardens, situations arise that must be dealt with and often quickly to maintain a healthy and productive environment. Plant cultivation at APC requires an investment to time to ensure the well-being of aquatic species in the collection. Phenological monitoring of species and their horticultural management. Fortnightly tours are carried out to identify the stages of development of the different species in the collection. The formation of flowers, fruits and/or seeds is mainly identified and recorded. These observations also help determine whether the specimens require pruning or transplantation. Comprehensive pest and disease management. Fortnightly observations also detect the presence of an insect or any other organism whose presence may cause some damage to aquatic species. Upon detection, Dr. Bonifacio Don Juan (responsible for the phytosanitary management of the collections) is notified and the recommended treatment is applied. Some problems caused by aphids, molluscs, and fungi have been detected. Their treatment has been timely and health problems have been kept under control (Figure 6). Wildlife program: native versus exotic species. The ponds provide habitat for wildlife. In some of the ponds and according to the season, you can observe various birds such as the Mexican duck (Anas platyrhynchos ssp. diazi), blue heron (Nycticorax nycticorax), frogs (e.g., Lithobathes montezumae), and dragonflies (e.g., Rhionaeschna multicolor). Other organisms flock to the ponds to get their prey (e.g., some insects become part of the diet of the collar lizard - Sceloporus torquatus) or to drink water and cool down like “tlacuache” (Didelphis virginiana). Unfortunately, however, there are also exotic wildlife such as the Japanese turtle (Traechemys scripta elegans) abandoned by visitors. Turtles are abandoned because their owners be306 Wetland Science & Practice October 2020

lieve that ponds are a shelter option for them, but they are not. The ponds do not have a warm and constant temperature for them and in many of them there is no food suitable for them. Also many of the turtles arrive sick and fight each other as there are many (more than 25 individuals) for a small pond (60 m2 approx.). Their presence poses a threat to the development of water lilies and the proper functioning of the system. The turtles consume leaves and buttons preferably of water lilies and climb the leaves and damage them. Since 2017, monitoring of ponds has been carried out for the detection of exotic aquatic organisms. In 2019, a study was conducted to assess the effect of the overpopulation of the Japanese turtle on one of the APC ponds. The turtle population was assessed, individual health was determined, then the turtles were removed from the pond, treated, and when they were in good condition they were put up for adoption so that they could be removed from the pond to evaluate the response of the plants by removing an element from the system. Four workshops have been held to inform visitors about the problem of abandonment of these turtles in the ponds. Incorporation of new species into exhibition ponds. Through links with other botanical gardens of the Mexican Association of Botanical Gardens A.C. specimens or seeds of aquatic plants have been received in donation for the strengthening of APC’s collection. The main donors have been the Botanical Garden of Fundación Xochitla, A.C. in the State of Mexico and the Botanical Garden of Culiacán, Sinaloa. Also some academics of the JB IB-UNAM, like biologist Ivonne Olalde, have collected aquatic plants that are currently propagated in the greenhouse for later incorporation into the ponds. In the last three years seven new species have been added to the collection: Anemopsis californica (Figure 7), Ludwigia sp., Thalia geniculata, Hydrocotyle ranunculoides, Limnobium laevigata, Marsilea mollis, and Lilaeopsis schaffneriana. Environmental education activities: the revaluation of Mexican aquatic plants. The general purpose of educational activities is to promote the public’s interest in aquatic plants and to recognize the importance of aquatic environments and their plant diversity. For four years (2015-2018) we participated in environmental education events: National Day of the Botanical Gardens and summer courses with the development of different activities related to the APC. More than 300 people have participated in the programs. Some activities were guided tours, children’s workshops or didactic games (Figure 8). The activities were based on the “Environmental Education Action Plan for the Botanical Gardens of Mexico” which promotes meaningful learning by involving different audiences and diverse strategies considering that it is not only important to know the biology of plants but to have a comprehensive vision of plant resources (Martínez-González et al. 2012). We have observed after the implementation of the activities that 90% of


the participants were unaware that there was so much diverFIGURE 8. Examples of environmental education activities to revalue sity of aquatic plants as they only knew the common water aquatic resources: guided tours (a) and children’s workshops (b). (Photos by APC – a, and Surya Ivonne González Jaramillo – b.) hyacinth (Eichornnia crassipes). They were most surprised to learn that some aquatic plants are so small (Lemna and Wolffia) while others are insectivorous (Utricularia). Participants also learned about the uses of aquatic plants since pre-Hispanic times and their relationship with aquatic fauna. Some suggestions for future activities include: 1) perform new activities considering different educational levels and age ranges, 2) implement workshops involving aquatic plant-animal relationships and 3) develop horticultural workshops for propagation of particular species such as horsetail (Equisetum hyemale). CHALLENGES: THE IMMEDIATE FUTURE APC will continue to develop the three basic lines of action: 1) sexual propagation of species, particularly for species of other botanical families or others with a different conservation status, 2) optimal development of species under ex situ conditions (e.g., assessing their growth and reducing the incidence of health problems), and 3) conducting environmental education activities and planning new strategies for educational intervention in locations (field sites) that still have natural populations of native aquatic vegetation. The Aquatic Plant Collection provides an experience that allows not only the study of Mexican aquatic plants, but is also like a blue and green island in an urban environment that allows the development of an artificial wetland with the presence of wildlife. In addition, it contributes to the development of new professionals (Figure 9) and offers Botanical Garden visitors an introduction to aquatic plants learning more about them (e.g. biology, ecology or ethnobotany) and the need for their conservation. n REFERENCES

Caballero Nieto, J. 2012. El Jardìn Botánico del Instituto de Biología y la Estrategia Global para la Conservación Vegetal. In: J. Caballero Nieto (ed.). Jardines Botánicos: contribución a la conservación vegetal de México. Ciudad de México, México: Comisión Nacional para el Conocimiento y Uso de la Biodiversidad. pp. 75-86.

FIGURE 9. Students in the greenhouse (Surya González, Diana Ferrusca, Nayeli González, Luis Silva, and Rogelio Yañez). (Photo by Surya Ivonne González Jaramillo.)

Bornette, G. and S. Puijalon. 2011. Response of aquatic plants to abiotic factors: a review. Aquat. Sci. 73: 1-14. López Martínez, C. 2018. Atlacuetzonan, planta acuática en el paisaje de la Cuenca de México a través de sus usos. Ciudad de México: Facultad de Arquitectura, UNAM. Lot, A. and A. Novelo. 2004. Iconografía y estudio de plantas acuáticas de la Ciudad de México. Ciudad de México. Universidad Nacional Autónoma de México. Martínez-González, L., and V. Franco, ad T. Balcázar. 2012. Plan de Educación Ambiental para los Jardines Botánicos de México. Ciudad de México. Asociación Mexicana de Jardines Botánicos. Miranda Arce, M.G. 1980. Plantas Acuáticas útiles del Valle de México. México, D.F.: Facultad de Ciencias, UNAM. Mora-Olivo, A., J.L. Villaseñor, and M. Martínez. 2013. Las plantas vasculares acuáticas estrictas y su conservación en México. Acta Botánica Mexicana 103: 27-63. Wetland Science & Practice October 2020 307


NOTES - WETLAND RESEARCH

Bark Traits: A Predictor for Recognition of Successional Groups in Riparian Forest Species Jane Rodrigues da Silva1,2, Augusto Cesar de Aquino Ribas3, Diogo da Silva Matos4, Edna Scremin-Dias2, and Rosani do Carmo de Oliveira Arruda2

F

rom the anatomical point of view, “bark” is the term used to designate all tissues exterior to the vascular cambium, therefore including peridermis, secondary phloem, and if they remain, the primary phloem and the cortex (Angyalossy et al. 2016). Bark traits, such as thickness and density, provide a remarkable data source that can be exploited to understand the plant’s life strategies in a community. Here, we evaluated whether the bark functional traits and anatomical features could be used as indicators of successional groups in woody species from the riparian forest of the Paraguay River, Mato Grosso do Sul state, Brazil. We hypothesized that pioneer species have thicker bark, with lower density and low percentage of sclerenchyma tissues than the late successional species. We collected bark samples from the main stem at 1.30 m above ground from three specimens in each group: pioneer species (Inga vera, Fabaceae; Triplaris gardneriana, Polygonaceae; Vochysia divergens, Vochysiaceae) and late successional species (Handroanthus heptaphyllus, Bignoniaceae; Ocotea diospyrifolia, Lauraceae; Vitex cymosa, Lamiaceae; Damasceno-Junior et al. 2005). Histological slides of the bark were prepared following standard techniques. We estimated the relative bark thickness (RBT) as the total bark thickness divided by bole diameter multiplied by 100 (Midgley and Lawes 2016). Bark density was calculated as dry mass per fresh volume (Borchert 1994). We estimated the area of sclerenchyma (sclereids and fibres cells; %) in cross-sections of the inner bark (secondary phloem and cortex, if any). We performed multidimensional scaling (MDS) analysis based on a Bray-Curtis dissimilarity matrix and used mixed models to test the difference between species groups using species as a random factor. A single periderm was observed in pioneer species (Figure 1a) and Handroanthus heptaphyllus, a late successional species. In this species, the phellem is formed by numerous layers of thick-walled, sclerified cells interspersed with few layers of thin-walled cell layers (Figure 1b). In Vochysia divergens (a pioneer species) the formation of aerenchymatous phellem (Figure 1d) was observed. In Vitex cymosa (Figure 1c) and 1 Correspondence author contact: janersbio@gmail.com 2 Laboratório de Anatomia Vegetal, Instituto de Biociências (INBIO), Universidade Federal de Mato Grosso do Sul, Campo Grande, Mato Grosso do Sul, Brazil. 3 Agência de Tecnologia da Informação e Comunicação, Universidade Federal de Mato Grosso do Sul, Campo Grande, Mato Grosso do Sul, Brazil. 4 Programa de Pós-Graduação em Biologia Vegetal, Universidade Federal de Mato Grosso do Sul, Campo Grande, Mato Grosso do Sul, Brazil.

308 Wetland Science & Practice October 2020

Ocotea diospyrifolia a thick rhytidome was observed. In pioneer species Triplaris gardneriana (Figure 1a) and Inga vera and, in late successional species Handroanthus heptaphyllus and Vitex cymosa the secondary phloem is formed by tangential layers of axial parenchyma and conducting cells alternating with tangential layers of sclerenchyma from vascular cambium towards periderm. The sclerenchyma is composed of large sclereids in Vitex cymosa, and fibers with thickened walls in other species. In late successional Ocotea diospyrifolia small aggregates of sclereids and large, oval sclereids clusters with thickened wall disperse were observed in the entire bark (Figure 1f). In Vochysia divergens, large sclereids clusters are restricted to the collapsed secondary phloem (Figure 1e). We detected starch stored in the parenchyma cells in all examined species and phenolic compounds in the inner bark of pioneer species Inga vera and Triplaris gardneriana and late successional species, Ocotea diospyrifolia. We observed that the bark traits differ between pioneer and late successional species (n = 18; t-value 2.33; p-value = 0.03), where RBT seems to be the major bark trait to determine the successional groups than the bark density (pioneer = 0.39 g/cm3; late successional = 0.40 g/ cm3) and percentage of sclerenchyma (pioneer = 30.8%; late successional = 32.6%) (Figure 1f). Contrary to our expectation, late successional species (4.61%) have thicker bark than pioneer species (2.99%). Thicker bark in late successional species can act in stem protection from decay and against pathogens and herbivores (Loehle 1988) which these species are prone to during periods of flooding in riparian forests. At the same time, thick bark consisting of rhytidome or phellem with numerous layers of thick-walled cells can play a key role in cambium protection from lethal temperatures during fires. In the face of the alarming increase of fire events in the riparian forest of the Pantanal, thick bark will be an important strategy and adaptive functional trait for woody species in these communities. Bark adaptation in pioneer species involves the development of aerenchymatous phellem in Vochysia divergens, a feature that can increase aeration in the bark tissues caused by anoxia during flooding (Yáñez-Espinosa and Terrazas 2001) and the accumulation of phenolic compounds in Inga vera and Triplaris gardneriana, which provides chemical defense in the bark (Roth 1981). Production of these compounds is stimulated by flooding in several species (Yule et al. 2018). n


REFERENCES

Angyalossy, V., M. Pace, R.F. Evert, C.R. Marcati, A.A. Oskolski, T. Terrazas, E. Kotina, F. Lens, S.C. Mazzoni-Viveiros, G. Angeles, S.R. Machado, A. Crivellaro, K.S. Rao, L. Junikka, N. Nicolaeva, and P. Baas. 2016. IAWA list of microscopic bark features. International Association of Wood Anatomists (IAWA) Journal 37: 517–615. doi:10.1163/22941932-20160151. Borchert, R. 1994. Soil and stem water storage determine phenology and distribution of tropical dry forest trees. Ecology 75(5): 1437–1449. doi:10.2307/1937467.

FIGURE 1. Examples of bark traits and ordination results: (a) General aspect of stem bark in transverse sections of the pioneer species Triplaris gardneriana. Note the layers of axial parenchyma and conducting cells alternating with layers of sclerenchyma (arrows) in the inner bark. (b) Phellem (Ph) with thickwalled, sclerified cells in Handroanthus heptaphyllus, a late successional species. (c) Rhytidome (Rh) with successive development of periderms (arrowhead) in a late successional species Vitex cymosa. (d) Aerenchymatous phellem in the pioneer species Vochysia divergens. (e) Secondary phloem in Vochysia divergens with large sclereids clusters (*). (f) Secondary phloem in a late successional species Ocotea diospyrifolia with small (arrows) and large (*) sclereids groups. (g) Multidimensional scaling (NMDS) ordination based on a Bray-Curtis matrix of dissimilarities showing the distribution of the pioneer and late successional species in the riparian forest of the Paraguay River. Scale bars: (a) = 500 µm; (b, c, d and e) = 100 µm; (f) = 100 µm.

Damasceno-Junior, G.A., J. Semir, F.A. Maës Dos Santos, and H. De Freitas Leitão-Filho. 2005. Structure, distribution of species and inundation in a riparian forest of Rio Paraguai, Pantanal, Brazil. Flora 200(2): 119–135. doi:10.1016/j. flora.2004.09.002. Midgley, J.J., and M.J. Lawes. 2016. Relative bark thickness: towards standardised measurement and analysis. Plant Ecology 217(6): 677–681. Springer Netherlands. doi:10.1007/s11258-016-0587-8. Roth, I. 1981. Structural Patterns of Tropical Barks. Gebrüder Borntraeger, Berlin. Yáñez-Espinosa, L. and T. Terrazas. 2001. Wood and bark anatomy variation of Annona glabra L. under flooding. Agrociencia 35(1): 51–63. Yule, C.M., Y.Y. Lim, and T.Y. Lim. 2018. Recycling of phenolic compounds in Borneo’s tropical peat swamp forests. Carbon Balance Management 13: 1–14.

Wetland Science & Practice October 2020 309


Mangrove Ecological Restoration Inside Pantanos de Centla Biosphere Reserve (Tabasco, México) Raúl Alejandro Betancourth1,2, Pilar A. Gómez-Ruiz3, and Paulo Carbajal-Borges2

antanos de Centla Biosphere Reserve (PCBR) is located in southern México within Tabasco state (Figure 1). It was declared as a Natural Protected Area by Federal Government in 1992 and in 1995 it was designated as a Wetland of International Importance (RAMSAR Convention Site 0733). PCBR is considered the most extensive and important wetland in Mesoamerica, being a strategic area due to the presence of migratory species and high biodiversity richness. Main ecosystems in PCBR are hydrophytic communities like popal-tular vegetation and mangroves represented by associations of Rhizophora mangle, Laguncularia racemosa and Avicennia germinans distributed in fringe and riverine mangrove types. Since PCBR is a protected area, a series of measures for conserving and protecting ecosystems biodiversity have been developed including wildlife monitoring, mangrove restoration, firewall breaches construction, forest watch and environmental education. However, natural protected areas are not exempted from impacts associated with global climate change, and for that reason PCBR managers recently started to implement a plan for mitigation and adaptation to climate change, to increase ecosystem resilience, to protect biodiversity, and to minimize vulnerabilities of local communities that inhabit this important region of southern México. Because of these needs, the “Resiliencia Project” has been developed to address this environmental problem in different protected areas of Mexico. Its main objective is to decrease direct and adverse impacts of climate change on biodiversity and human communities through strengthening the effectiveness of management and spatial configuration of México´s Natural Protected Areas. This project has been implemented in 17 Natural Protected Areas in México, including PCBR. The initiative was funded by the Global Environmental Fund (GEF) and was implemented by United Nations Development Program (UNDP) and National Commission of Natural Protected Areas (CONANP in Spanish). Project implementation in PCBR was conducted between 2019 and 2020 by the Mexican NGO - Foro para El Desarrollo Sustentable A.C., with academic and technical support from

P

1 Corresponding author contact: rabetancourthb@gmail.com. 2 Foro para el Desarrollo Sustentable A.C., San Cristóbal de Las Casas, Chiapas, México. 3 CONACYT-Universidad Autónoma del Carmen, Campeche, México.

310 Wetland Science & Practice October 2020

the Universidad Autónoma del Carmen and Biomasa A.C NGO. The main actions executed during the project were: • Evaluation of socio-ecological conditions to determine the technical viability and necessity of mangrove ecological restoration. • Implementation of ecological restoration actions on 50 hectares through reforestation and hydrologic rehabilitation of mangroves. • Conducted training workshops to strengthen local capacities for conserving and managing mangroves, with gender inclusion as a strategy for climate change adaptation and extreme weather phenomena risk prevention. The project was implemented at El Palmar and Tembladeras ejidos, both located on northern PCBR (Figure 1). El Palmar community has 87 people and Tembladeras has 122 people. These are communities that have communal property control over land, and thereby make collective decisions via an assembly of communal landowners (ejidatarios). Their main economic activity is fishing for self-consumption and selling in local markets. PARTICIPATORY ECOLOGICAL RESTORATION PROCESS Project development was done through a participatory framework, where local knowledge and experience play a fundamental role. Every decision related to actions, design, techniques, and work seasons was made after dialogue and coordination among communities, local authorities and the PCBR managers. A diagnostic evaluation was carried out with socioecosystemic focus to evaluate the viability of different sites for restoration, both in El Palmar and Tembladeras. To do this, we used three sources of information: 1) cartographic analysis, 2) gathering of local knowledge, and 3) ecological evaluation. Through cartographic analysis, we found vegetation distribution patterns that were helpful in locating biological corridors. We were also able to correlate geography with historical data on wildfires and obtain local knowledge about fire-prone areas that helped us detect and avoid sites that are under constant threat from fire. According to the analysis of all this information, the most suitable restoration action for ejido El Palmar was


mangrove reforestation with Rhizophora mangle propafires. This participatory ecological restoration process aims, gules (Figure 2). On the other hand, for ejido Tembladeras in the short- and long-terms, to recover mangrove ecosysthe needed action was hydrologic rehabilitation through tem at landscape level and to increase local resilience and cleaning natural channels in a previously reforested area improve climate change adaptation by local communities where mangroves grew up without control and caused hyinside PCBR. n drologic connectivity problems at the landscape level. A few months FIGURE 1. Pantanos de Centla Biosphere Reserve (PCBR) and case study area localization. (A) General later, during the first monitoring overview of Mexico with study area indicated, (B) state of Tabasco, indicating PCBR location, (C) PCBR area with case study locations “ejidos” El Palmar and Tembladeras and (D) Amplification of both ejidos surface. evaluation of restoration activities, we recorded a global survival Elaborated by Juan Paulo Carbajal Borges based on Instituto Nacional de Estadística y Geografía (INEGI). 2010. Red Carretera de México, Comisión Nacional de Áreas Naturales Protegidas (CONANP) 2017. Áreas of 75% of R. mangle propagules Naturales Protegidas Federales, Registro Agrario Nacional (RAN) 2017. Perimetrales de Núcleos Agrarios. planted in ejido El Palmar. In the case of Tembladeras, fish species of interest were captured in rehabilitated channels, providing evidence that hydrologic connectivity is recovering. Training workshops were planned to cover these topics: mangrove ecosystem services, methodological phases of mangrove ecological restoration projects, and restoration monitoring techniques. In addition, a basic course on forest fires was given, with the approval of the National Forest Commission (CONAFOR in Spanish). A community-experiences exchange was done with a neighborhood community that had wide experience in mangrove sustainable use and restoration practices in Tabasco state. In this FIGURE 2. Hydrophytic vegetation in channels or “picadas” where reforestation action was implemented in activity, people from El Palmar ejido El Palmar, PCBR. (Photo by Alejandro Betancourth.) and Tembladeras communities visited a mangrove nursery, restored sites with L. racemosa, and a charcoal production operation. Currently, Foro para el Desarrollo Sustentable A.C. together with its academic and technical allies, are continuing the initiative through a variety of activities including additional restoration actions, developing a protection strategy for restored sites through establishment of firewall breaches, designing local climate adaptation and disaster risk reduction plans, and implementing early warning systems for mangrove Wetland Science & Practice October 2020 311


Diversity of Temporary Ponds from the Guajira, Colombia Cesar E. Tamaris-Turizo1, Pedro Eslava, Juan M. Fuentes-Reinés, Daniel Serna-Macías, Diana Tamaris, and Luis Castro

emporary ponds are ecosystems of great importance for the biodiversity of the arid region of northern Colombia. These temporal systems support resident and migratory animals including birds, frogs, snakes, plankton, macroinvertebrates, and fishes. These organisms display strategies, for example, some invertebrates burrow their larvae in the sediment or produce resistant eggs in diapause state, that permit them to resist the drought periods that could last more than six months. Despite their importance to local and migratory species, temporary ponds in Colombia have received little attention from researchers. The aim of our study was to characterize the biodiversity of five temporary ponds located on the middle Guajira (Maicao, Manaure, Uribia, Mayapo and El Ebanal) of northern Colombia and compare the communities between sites (Figure 1). Examples of these ponds are shown in Figure 2.

Biological samples were taken during wet period of 2018. The plankton samples were collected using a 25-l bucket at both littoral and limnetic habitats. Samples were filtered with a zooplankton net (mesh size = 45 μm) and phytoplankton net (mesh size = 23 μm) and preserved in 70% ethanol. Macroinvertebrates associated with macrophytes were taken with a hand net (mesh size = 300 μm) and macroinvertebrates associate with sediment with an Ekman dredge. The fishes were taken by tie and hand net (both, mesh size = 500 μm), some fishes were transported to laboratory to confirm the species. Birds were identified though free tours around the ponds from 5:00 am to 9:00 am and from 4:00pm to 6:00pm. The fauna found in the systems of temporary ponds of La Guajira contain typical tropical species. More than 12,000 organisms were collected and observed, belonging to five large groups: macroinvertebrates, microphytes, fishes, birds, and zooplankton). A total of 10,465 macroinvertebratres were collected, representing 30 genera FIGURE 1. General locations of the temporary ponds (The base image was taken from https://www. and 20 families. Among this group, laguajira.gov.co/web/la-guajira/division-politica-administrativa.html). Triops (tadpole shrimp), Dendrocephalus (fairy shrimp), and Thamnocephalus (fairy shrimp) were the most abundant genera with 3,207, 2,269 and 3,169 individuals tabulated, respectively. Triops and Thamnocephalus are first records from La Guajira. A total of 204 taxa of microphytes were identified, grouped in 66 genera and 40 families. Euglenophyceae was the family with greatest richness (25% of all families) whereas Trachelomonas and Lepocinclis were the most abundant genera of euglenoids. Only seven species of fishes were found Austrofundulus guajira (killifish), Rachovia hummelincki (killifish), Astyanax magdalenae (tolomba), Mugil incilis (mullet), Dormitator maculatus (whitebait), Eleotris amblyopsis, and Ctenolucius hujeta (hujeta gar). The two

T

1 Grupo de Investigación en Ecología y Biodiversidad, Universidad del Magdalena, Santa Marta, Colombia; corresponding author contact: ctamaris@unimagdalena.edu.co.

312 Wetland Science & Practice October 2020


former species occurred in isolated ponds of fluvial systems while the others were found in ponds with influence of lotic ecosystems. Fifty-two bird species were observed, with the Carib grackle (Quiscalus lugubris) being most frequent. The zooplanktonic community were represented by eight taxa of Rotifera (Lecane sp, Platyas brevicornis, Asplanchna sp, Brachionus caudatus, B. quadridentatus quadridentatus, B. mirabilis, Filinia sp., and Dicranophorus sp), 18 of Cladocera (Diaphanosoma breverime, Diaphanosoma spinulosum, Diaphanosoma dentatum, Sarsilotona serricauda, Pseudosida sp., Ceriodaphnia cornuta, Moinodaphnia macleayi, Moina micrura micrura, Moina reticulata, Macrothrix elegans, Macrothrix spinosa, Grimaldina freyi, Kurzia polyspina, Leydigia cf striata, Ovalona cf glabra, Chydorus nitidulus, Dunhevedia odontoplax, and Dunhevedia crassa) and four taxa of Copepoda (Prionodiaptomus colombiensis, Microcyclops cf ceibaensis. Mesocyclops brasilianus, and Thermocyclops tenuis). Of these Platyias quadricornis quadricornis, Brachionus caudatus, B. quadridentatus quadridentatus, B. mirabilis, Prionodiaptumus colombiensis, Mesocyclops brasilianus, and Thermocyclops tenuis are new records to the Department of La Guajira. The most abundant zooplankton species was a copepod - Prionodiaptomus colombiensis (860 individuals) whereas the least abundant was a rotifer - Asplanchna sp. with only two individuals collected. Lecane sp., Thermocyclops tenuis, Microcyclops cf ceibaensis, Moina micrura, Moinodaphnia macleayi, Diaphanosoma spinolusum, and Leydiga cf striata were the most frequent taxa. The temporary pond with the greatest richness of zooplankton was Maicao pond with 20 species, while El Ebanal pond had the least richness with only nine species. Overall, Maicao ponds had the highest biodiversity, probably due to the habitat heterogeneity marked by the occurrence of different patches of macrophytes (i.e., Eleocharis elegans and Paspalum sp.) and scrubs (i.e., Caesalpinia coriaria and Parkinsonia praecox). Research about the community structure in temporal ponds in Colombia is scarce and it is expected that the results of our study will help stimulate similar investigations in order to generate local listings that reveal new records of species occurrence and, consequently, to better understand the diversity of the fauna of Colombia’s temporary ponds. n

FIGURE 2. Sampled temporary ponds in the northern Colombia: a) Maicao, b) Uribia, c) Manaure, d) Mayapo, and e) El Ebanal. (Photos by P. Eslava). Note: The first four ponds (a, b, c and d) are temporary ponds isolated from fluvial systems, while the last pond (e) is connected with the RancherĂ­a River.

Wetland Science & Practice October 2020 313


Treatment Wetlands - Experiencies in Mexico Armando Rivas1

I

n Mexico treatment wetlands are mostly utilized for wastewater treatment, mainly due to their low operating costs, simple operation, aesthetics and successful experiences. Approximately 40% of the treatment infrastructure (mainly systems requiring electricity) are out of operation, due to, among other causes, insufficient economic resources and slightly trained personnel. There is therefore a need for new technological alternatives to help solve this serious problem. In the Patzcuaro Lake basin, the installed treatment capacity (activated sludge and oxidation ditches) is 190 L/s, of which 52 L/s are treated (72% of the flow rate was used for agricultural activities before getting to treatment plant). There is also the problem of discharging residual sludge. The problem is compounded by the direct discharge, into the lake, of wastewater generated in agricultural and industrial activities, which cause eutrophication into the lake. An additional problem, perhaps the most important is concerned to social considerations, since there is an obvious rejection of conventional systems for the above mentioned reasons and because people don’t like the bad odors generated from wastewater treatment facilities. As an alternative solution and through the implementation of an intensive social participation program (workshops, videos, brochures, visits to wetlands in operation, legal aspects of land, etc.) treatment wetlands were

installed in several communities. Each system is composed of: pretreatment (screening and grit removal), septic tank, vertical subsurface flow wetland (sludge wetland planted with Arundo donax), subsurface flow wetland (planted with Typha sp.), and maturation pond and subsurface flow wetland (with Typha sp.). Treatment wetlands were created in five communities (2004 in Cucuchucho, 2005 in Santa Fe de la Laguna, 2006 Erongaricuaro, 2007 San Francisco Uricho and two more, 2007, at San Jeronimo Purenchecuaro. So by 2007, a total of six wetland treatment systems had been built in the Patzcuaro Lake basin. Prior to the construction of the systems, studies of wastewater characteristics, topography and soils mechanics were carried out. After one year of operation, the first system (Cucuchucho locality, for 1.0 L/s) was evaluated during the dry season (December to May) by four sampling campaigns. Average influent and effluent values (and removal efficiencies –%-) were: BOD 454 and 12 mg/L (97%); TN 79 and 10 mg/L (87%); TP 17 and 6 mg/L (65%); FC 3.42E 06 and 5.72E 02 MPN/100 mL (99.9834%), and helminths 12 and 0 (100%). The Mexican regulation (NOM-001-SEMARNAT-1996, discharge into lakes and protection of aquatic life) are: BOD 30 mg/L, TN 15 mg/L; TP 5 mg/L, Helminths 1/L and FC 1000 MPN/100 mL. So for this treatment all parameters were met except for the TP standard. Figure 1 shows the BOD

FIGURE 1. Results for Cucuchucho wetland treatment system during first year of operation. BOD

1,000

Constituents, mg/L

TN TP

100

10

1

Inlet

1 Instituto Mexicano de Tecnología del Agua. Av. Paseo Cuauhnáhuac Nº 8532 Col. Progreso. Jiutepec. Morelos. México. C.P. 62550; rivas.hz@gmail.com

314 Wetland Science & Practice October 2020

Septic Tank

Wetland 1

Pond

Wetland 2


results (graph on left) during the four sampling campaigns (dry season) and the average results of BOD, TN, and TP for each component of the system (graph on right). One year later, a second system (3.0 L/s) was built in the community of Santa Fe de la Laguna. The obtained efficiencies were similar to those achieved in the Cucuchucho system. The impact of treated water on lake water quality was subsequently assessed by a study of bio indicators (macroinvertebrates), used in their concept of Family Level Biotic Index (FBI), in four sampled points (locations with or without treatment wetlands). After two years of the start-up of Cucuchucho system (first evaluated point), the FBI was 12; at the second point (Santa Fe de la Laguna, a system operating after one year of the start-up), the FBI was 10; the third point, near Opongio community (blank site, inside the lake, with no treatment plant or wastewater discharge) the FBI was 9; and the fourth place (Tzintzuntzan community, without treatment plant, therefore direct discharge of waste water into the lake) the FBI was 0. The BOD, inside the lake, near the discharge of treated water, before the installation of the treatment wetland, was 414 mg/L. After three years of water treatment by the wetland system BOD dropped to 88.9 mg/L, indicating a significant improvement in water quality within the lake from wastewater treatment by wetlands. Other benefits by using wetlands were: 1) sale of ornate flowers (Zantedeschia aethiopica, placed on the periphery to improve aesthetics), 2) sale of Typha angustifolia as raw material for handicrafts manufacturing, 3) treated water that complies with the regulations for crops irrigation (vegetables and

maize), and 4) areas to protect wildlife (lake birds). The successful application of wetland systems for wastewater treatment now serves as a model for other communities (total of five built systems) and have been used for education on the value of natural treatment systems in workshops of water culture for several communities. In addition, the wetlands are frequently visited by schoolchildren, academics, and people interested in the use of this technology for wastewater treatment. After installation of the five systems, water quality studies were carried out at the sites where the treated water is discharged, either through wetlands or with electromechanical systems. The results, after ten years of operation, indicate there is a lower concentration of nutrients (TN and TP) in the lake where wetlands and lagoons are functioning, in comparison with where the electromechanical systems are employed. It was also observed that the efficiency of TP removal by conventional treatment plants is decreasing after the start-up; a cause, which is likely to contribute to this result, could be due to insufficient activities of operation and maintenance. In conclusion, the use of treatment wetlands has produced successful results by improving the quality of lake’s water, protecting the lake’s biodiversity and health of local people and livestock, providing additional income through the sale of by-products, and creating areas for wild life protection. Perhaps best of all, the wetland treatment systems keep on working with a minimum of attention and have been widely accepted in social terms by local communities. n

Wetland Science & Practice October 2020 315


NOTES - WETLAND CONSERVATION/EDUCATION/OUTREACH

An Ecosystem-based Approach to Managing Fish, Cattle and Forests on the Amazon Floodplain David G. McGrath1,2, Antonia Socorro Pena da Gama2,5, Laura L. Hess3, Bruce Forsberg4, Antonio José Mota Bentes5, Poliane Batista5

he Lower Amazon floodplain is the vast fluvial wetland associated with the Amazon River upstream of its internal delta and estuary. It extends from above the mouth of the Xingu River in the East to the border between the states of Pará and Amazonas in the West and has a total area of 16,000 km2, of which roughly 11,000 km2 is seasonally inundated (Figure 1). It is characterized by large open water bodies surrounded by seasonally flooded grasslands and forests that vary in extent and inundation state during the annual flood cycle. Forests occupy the levees bordering river channels while grasslands occupy the transition zone between the levees and the open water bodies of the floodplain interior (Figure 2). The river level peaks in early June, falls to its minimum level in early November and then rises gradually from December to June. Flood amplitude is about 6 meters, with the highest floods exceeding 8 meters. At peak high-water level approximately 95% of the floodplain is inundated, with open water, flooded grasses and flooded forests occupying 59, 21 and 20% of this area, respectively (Hess et al. 2003). At low water, only 67% of the floodplain is flooded, with open water, flooded grasses and flooded forests occupying 69, 18 and 13% of this flooded area, respectively (Figures 3 and 4). Many trees have adapted to this flood regime by producing fruits and seeds that mature as floodwaters reach their peak, providing an important food source for fish, which in turn act as a key dispersal agent for the plants (Goulding et al. 2019; Schöngart et al. 2002). Many species of fish, including important commercial species such as Characins, use the floodplain as nursery habitat and – once mature – as seasonal feeding habitat. As adults these species move out of the floodplain to migrate upstream to spawn during rising water levels and then move back onto the floodplain to feed during the high-water period. The Lower Amazon is the most deforested section of the Amazon floodplain. During the 18th and 19th Centuries, levee forests were cleared for cacao plantations. Jute was

T

1 Earth Innovation Institute, San Francisco, CA, USA; 2 Federal University of Western Pará, Santarém, PA, Brazil; 3 Earth Research Institute, University of California , Santa Barbara, CA, USA; 4 Vermont Department of Environmental Conservation, Montpelier, VT, USA; 5 Sapopema (Sociedade para a Pesquisa e Proteção do Meio Ambiente), Santarém, PA, Brazil.

316 Wetland Science & Practice October 2020

introduced by Japanese colonists in the 1930s and after WWII, it expanded throughout the Amazon floodplain, reaching a peak in the 1970s before declining through the mid-1980s when production ended. Since the colonial period, the natural grasslands of the floodplain have been used to graze cattle. This intensified in the 1980s with the seasonal movement of cattle between floodplain and uplands. Today, cattle ranching and fishing are the dominant resource use activities on the floodplain. Between the mid1970s and 2008 more than 50% of forest cover was cleared and today there is no primary forest on the Lower Amazon floodplain (Renó et al. 2016). Cattle ranching is practiced on a range of scales from artisanal fishers who raise a few head of cattle to largescale ranchers with herds of several hundred. Deforestation and degradation of forests and grasslands is, in part, the result of the economic strategies of small-scale fishers who in addition to fishing in floodplain lakes, practice shifting cultivation for annual crops and raise cattle on community grasslands (Figure 2). Within household economic strategies, fishing provides subsistence and cash income, while cattle serve as a source of savings. While small-scale fishers understand the close relationship between floodplain fisheries and forests, many continue to raise cattle, undermining management efforts to increase the productivity of local fisheries. The tendency for smallholders to invest in cattle as a savings strategy, despite the consequences for the local fisheries on which they depend for their cash income, is related to the perceived security of investments in cattle versus those in managing the fishery. Insecurity regarding fisheries is exacerbated by the lack of government enforcement of management agreements. Property rights to fish in a lake are diffuse and follow the law of capture, while property rights to cattle grazing on collectively owned grasslands are clear and enforced by the state. Consequently, smallholders invest in fishing for immediate needs (subsistence and cash), while for long-term savings, they invest in cattle. Until these conditions change, it is unlikely that smallholders will agree to reduce their investment in cattle to conserve the habitat that improves fisheries productivity.


To address the problems of FIGURE 1. Location map showing the Lower Amazon region centered on the city of SantarĂŠm, at the overfishing in community lakes confluence of the TapajĂłs and Amazon Rivers. The box shows the area covered in Figure 3. and overgrazing of floodplain grasslands, communities have implemented inter-community fishing agreements and formal contracts between individual communities and local cattle owners. FIGURE 2. Transect of the Lower Amazon floodplain showing main habitat types and resource use by These informal fishing agreements floodplain smallholders. led to the development of a comanagement policy that provides a pathway for their legal recognition. Communities then adapted this model to negotiate agreements between communities and local cattle owners to regulate cattle grazing on floodplain grasslands. These were legalized through the Public Ministry via a contract called a Term of Adjustment of Conduct (TAC). Some 50 TACs FIGURE 3. Vegetation cover of Lower Amazon floodplain at low-water stage (top) and high-water stage were negotiated by communities (bottom). in the region. Between 2006 and 2008, community territories were incorporated into 41 Agro-extractive Settlement Projects (PAE), a type of extractive reserve for traditional populations, with a total area of 740,000 hectares. Both agreements were incorporated into the Utilization Plans of the PAE, creating the potential for the integrated management of settlement resources (Figure 5). However, lack of enforcement continues to 2018). These studies make it possible to calculate the cost be a problem. to fishers of forest and grassland degradation caused by Floodplain residents understand the importance of cattle and shifting agriculture. These economic estimates floodplain forests and grasslands to their livelihoods, not can then be used to generate scenarios so fishers can clearly only as feeding and nursery habitat for fish, but also for see the costs of investing in cattle when designing lake controlling bank erosion and as wave barriers to protect management plans that regulate fishing and cattle grazing houses during flood season thunderstorms. There have been and reforestation of floodplain lake margins. We plan to numerous initiatives to reforest levees and replant aquatic use this approach in working with floodplain communimacrophytes, but these are often frustrated by uncontrolled ties to negotiate more effective agreements for regulating cattle grazing. cattle and to justify community investments in large-scale In recent years, several research initiatives have docureforestation projects that improve the productivity of lake mented the close relationship between the extent of forest fisheries and increase aquatic biodiversity and the resilience cover surrounding lakes and the productivity of fishing of floodplain fisheries and household subsistence strategies. effort at both the regional scale and through comparison Our strategy going forward builds on these community of fishing productivity, fish diversity and fish commuinitiatives for reforesting lake margins, planting floating nity structure in individual floodplain lakes with varying grass storm barriers and other local approaches to conservamounts of forest cover (Castello et al. 2018; Arantes et al. ing habitat. This partnership harnesses the potential of Wetland Science & Practice October 2020 317


universities in Santarém through collaborations with local NGOs, community schools and community management organizations, combining science, local knowledge and local organizational capacity to sustainably manage floodplain fisheries, grasslands and forests. n REFERENCES:

Arantes, C. C., K.O. Winemiller, M. Petrere, L. Castello, L.L. Hess, and C.E.C. Freitas. 2018. Relationships between forest cover and fish diversity in the Amazon River floodplain. Journal of Applied Ecology 55: 386–395. Castello, L., L. L. Hess, R. Thapa, D. G. McGrath, C. C. Arantes, V. F. Renó, and V. J. Isaac. 2018. Fishery yields vary with land cover on the Amazon River floodplain. Fish and Fisheries 19(3): 431-440. Goulding, M., E. Venticinque, M. L. de B. Ribeiro, R. B. Barthem,R. G. Leite, B. Forsberg, P. Petry, U. Lopes da Silva-Júnior, P. Santos Ferraz, and C. Cañas. 2019. Ecosystem-based management of Amazon fisheries and wetlands. Fish and Fisheries 20(1): 138-158.

Hess, L. L., J. M. Melack, E. M.L.M. Novo, C. C.F. Barbosa, and M. Gastil. 2003. Dual-season mapping of wetland inundation and vegetation for the central Amazon basin. Remote Sensing of Environment 87: 404–428. McGrath, D. and M. Crossa. 1998. Restoration of floodplain lake habitat: a PLEC demonstration project. PLEC News and Views 11: 10-17. McGrath, D., A. Cardoso, O.T. Almeida, and J. Pezzuti. 2008. Constructing a policy and institutional framework for an ecosystem-based approach to managing the lower Amazon floodplain. Environment, Development and Sustainability 10: 677-695. Renó, V., E. Novo, and M. Escada. 2016. Forest fragmentation in the Lower Amazon Floodplain: Implications for biodiversity and ecosystem service provision to riverine populations. Remote Sensing 8: 26. Schöngart, J., M. T. F. Piedade, S. Ludwigshausen, V. Horna, and M. Worbes. 2002. Phenology and stem-growth periodicity of tree species in Amazonian floodplain forests. Journal of Tropical Ecology 18: 581–597.

FIGURE 4. Contrasting low-water (left) and high-water (right) conditions, Curuai Lake, Amazon floodplain. PlanetScope imagery for 2 November 2017 (low) and 29 July 2017 (high). Images courtesy of Planet Labs.

FIGURE 5. Distribution of Agro-extractive Settlement Projects (PAE) of the Lower Amazon floodplain.

318 Wetland Science & Practice October 2020


Strengthening Governance in the Monterrico Multiple Use Natural Reserve: Planning for Conservation with a Bottom-Up Approach Ana Silvia Morales1

he Monterrico Multiple Use Natural Reserve (MMUNR) is an important protected wetland with high conservation values located on the Pacific Coast of Guatemala (Figure 1). It is comprised of estuarine and coastal-marine ecosystems that provide goods and services to the local people whose livelihoods depend on them, such as fisheries, tourism, coastal protection, among others. Since the MMUNR was created, like most of the protected areas back in the 1970s, there was no consultation process with local communities to obtain prior and informed consent to be part of conservation effort. The reserve is managed by the Center of Conservation Studies (CECON) of the University of San Carlos, a research center that also manages six other protected areas. Their objectives are to conduct permanent biodiversity research programs aimed

to improve the current biodiversity and ecosystem services management by creating new and pertinent models for protected areas management. In this regard, after an assessment of the current wetland management approach, it was necessary to redesign the way CECON was interacting with the local communities to achieve better conservation outcomes (Figure 2). Taking into account that the MMUNR is a wetland of considerable relevance for local people, CECON is working on a governance platform through updating its master plan with a bottom–up approach to include inputs from local communities regarding their perceptions and concerns about conservation of ecosystem goods and services for reducing threats to the MMUNR wetland, and also to get local people to interact with public and private sectors. This updating process is also considerFIGURE 1. Location of the MMUNR in the Pacific Coastal of Guatemala, Central America (Miguel à vila, ing actions beyond the borders of CECON). the protected area, by focusing on activities related to watershed management taking place upstream that are altering the natural hydrological cycle of the wetland. The main goals of this participatory process are: 1) to strengthen CECON as the protected area manager with the continued support of the local communities and the other stakeholders for decision making, and 2) establishing a committee to create an environment for sustainable use of the wetland ecosystems through a mechanism of effective and equitable governance that includes all sectors involved in the area. This process also contemplates an expansion of the role of CECON as the MMUNR Manager to include serving as both technical–scientific advisor and mediator for conflicts related to the use of

T

1 Coordinator of the Monterrico Multiple Use Natural Reserve, Center of Conservation Studies, University of San Carlos of Guatemala; ansilmo@gmail.com Wetland Science & Practice October 2020 319


biodiversity and natural resources. Moreover, recent international cooperation has been directed to strengthen conservation of coastal marine protected areas. In this regard, CECON is helping local communities implement productive sustainable practices in agriculture, strengthening local community organization, and directing conservation incentives towards those stakeholders involved in conservation and restoration of the protected area. The governance process carried out so far has achieved great success in terms of community involvement as local people have become more aware of the relevance of working together for the conservation of biodiversity and natural

resources. CECON is recognized as a management authority not only regarding the protected area but in general for environmental concerns in other local communities outside the borders. As a result of this ongoing shared work, a mutual collaboration agreement has been signed by both CECON and the local communities. The main challenge is to consolidate this process with the active participation and consent of the local communities together with the public and private sectors. In time, this can be achieved through the equitable sharing of benefits from the wetland conservation for all stakeholders in the area. n

FIGURE 2. Guatemala’s Monterrico Multiple Use Natural Reserve is important to both people and wildlife. (Photos by Ana Silvia Morales and Homero Escobar.)

320 Wetland Science & Practice October 2020


Amphibious Colombia: A Country of Wetlands Ronald Ayazo Toscano1, Wilson Ramirez, Ana Carolina Santos, Olga Lucía Hernández-Manrique, Angélica Batista, and Margarita Roa

he Biological Research FIGURE 1. Map of Colombia and its amphibious nature prepared by César Aponte adapted from Amphibian Colombia. A country of wetlands (Jaramillo Villa et al. 2015). This map shows the general distribution of five Institute Alexander von wetland categories in the five hydrographic areas of Colombia: Amazon, Caribbean, Magdalena-Cauca, Orinoco, Humboldt is a civil nonand Pacific. (Source: The Biological Research Institute Alexander von Humboldt, Bogotá.) profit organization associated with Colombia’s Ministry of Environment and Sustainable Development (MinAmbiente). It is part of the Colombian National Environmental System (SINA), which aims to store and share data to generate knowledge about Colombian environment and ecosystems. The von Humboldt Institute is recognized internationally as an authority on the status and trends of biodiversity, invasive species, biological collection, open data, citizen science, ecological restoration, ecohydrology, and supportive knowledge networks. The Institute conducts scientific research on biodiversity in the continental territory of Colombia, including freshwater resources. The ous Colombia: A country of wetlands, the most complete von Humboldt Institute has been an active participant in supporting MinAmbiente in the construction of Colombia’s and extensive scientific work about Colombian’s wetlands (Jaramillo Villa et al. 2015). This project created figures and National Wetland Policy (Naranjo et al. 1999). Colombia has a heterogeneous topography and geogra- images to guide the natural resource management and risk management of wetlands in the country, beginning with a phy that requires continuous study and up-to-date cartogcarefully developed set of maps (Figure 1). raphy to inform environmental policy-making. In contrast One of those maps, using official information later pubto other countries of the Andes region, Colombia has three lished by Flórez-Ayala et al. (2016), showed more than 30 distinct mountain ranges separated by rivers, creating a million hectares of wetlands, encompassing 26% of the nadiverse hydrographic network supporting a vast wetland tional inland area (see figure above). Wetlands were classiecosystems. To support its work, the von Humboldt Instified into four categories: open permanent, permanent under tute has several institutional associations, such as the Nacanopy, temporary, and potential. Open permanent repretional Adaptation Fund who financed the project Amphibisents wetlands where water presence is constant and there is no tree cover (e.g., lakes, lagoons, ciénagas, rivers, and 1 Instituto de Investigación de Recursos Biológicos Alexander von Humboldt, glaciers). Permanent under canopy identifies forested wet-

T

Bogotá, Colombia; Corresponding author: rayazo@humboldt.org.co.

Wetland Science & Practice October 2020 321


lands that are always flooded (e.g., Atrato River swamps or mangroves). Temporary wetlands do not have water year-round, but are periodically wet. The last category - Potential - identifies areas where soil and/or landform characteristics indicate the likely presence of a wetland, although no flooding was detected during the analyzed period (2007-2011); this category must be examined at scales with greater detail. Another map of interest identified and classified 89 macro-habitats across marine-coastal, inland, and anthropogenic systems using the Brazilian wetland classification system with geomorphological adjustments for Colombia (Ricaute et al. 2019). Lastly, in association with Ecopetrol (the Colombian State Petroleum Company), the von Humboldt Institute identified 28 freshwater ecoregions for Colombia, based on hydro-geographic and biological criteria like fish species composition, the interpretation of their drainage network, and geomorphological characteristics (Mesa-S et al. 2016). All these findings are provided to help improve and facilitate the creation of laws and public policies that include the protection, restoration and sustainable use of freshwater ecosystems necessary for the territorial planning. n

322 Wetland Science & Practice October 2020

REFERENCES

Flórez-Ayala, C., L.M. Estupiñán-Suárez, S. Rojas, C. Aponte, M. Quiñones, O. Acevedo, S.P. Vilardy Quiroga, and Ú. Jaramillo Villa. 2016. Identification and mapping of Colombian inland wetlands. Biota Colombiana 17(1):179–207. https://doi.org/10.21068/c2016s01a03 Jaramillo Villa, Ú., J. Cortés-Duque, and C. Flórez-Ayala (Editors). 2015. Amphibian Colombia. A country of wetlands. Alexander von Humboldt Research Institute of Biological Resources. Bogotá D.C., Colombia. http://hdl.handle.net/20.500.11761/34868 Mesa-S., L.M., G. Corzo, O. L. Hernández-Manrique, C.A. Lasso, and G. Galvis. 2017. Freshwater ecoregions from Colombia: a proposal for territorial planning of the Trasandean region and part of the Orinoco and Amazon basins. In: L.A. Moreno, G.I. Andrade, and L.F. Ruíz-Contreras (Editors). 2016. Biodiversity 2016. Status and Trends of Colombian Continental Biodiversity. Research Institute of Biological Resources Alexander von Humboldt. Bogotá D.C., Colombia. http://reporte.humboldt.org.co/biodiversidad/en/2016/cap4/406/index.html Naranjo, L., G. Andrade, and E. Ponce de León. 1999. Humedales interiores de Colombia: Bases técnicas para su conservación y uso sostenible.: Instituto de Investigación de Recursos Biológicos Alexander von Humboldt y Ministerio del Medio Ambiente, Bogotá D.C., Colombia. Ricaurte, L.F., J.E. Patiño, D.F.R. Zambrano, J.C. Arias-G, O. Acevedo, C. Aponte, R. Medina, M. González, S. Rojas, C. Flórez, L.M Estupinan-Suarez, U. Jaramillo, A.C. Santos, C.A. Lasso, A.A. Duque, S.R. Calle, J.I. Vélez, J.H. Caballero, S.R. Duque, M. Núñez-Avellaneda, I.D. Correa, J.A. Rodríguez-Rodríguez, S. Vilardy, A Prieto-C, A. Rudas-Ll, A.M. Cleef, C.M. Finlayson, and W.J. Junk. 2019. A classification system for Colombian wetlands: an essential step forward in open environmental policy-making. Wetlands 39: 971–990. https://doi. org/10.1007/s13157-019-01149-8


Amphibian Territories in Transition: Socio-ecological Rehabilitation of Wetlands Ronald Ayazo Toscano1, Wilson Ramírez1, Klaudia Cárdenas1, Olga Lucía Hernandez-Manrique1, Natalia Gómez López2, William Vargas2, Paola Isaacs-Cubides1, Mauricio Aguilar1, Yenifer Herrera1, Henry Huertas1, Juan C. Linares3, Wendy López4, and Jorge Bedoya5

T

he successful rehabilitation of a wetland landscape requires more than a single restoration method, especially when productive activities are carried out there. In Colombia, the La Niña phenomenon from 2010-2011 was particularly destructive, causing significant damage to infrastructure, economy, and human lives. For this reason, in The Mojana, a process of rehabilitating wetlands through socio-ecological services and ecosystem restoration was initiated to increase the adaptive capacity of natural ecosystems and the livelihoods of people dependent on it by reducing the risk of flooding and drought associated with climate change and its variability. The Mojana Region is an internal delta located at the south center of the Colombian Caribbean region (Figure 1). It is a floodplain landscape with marshes, swamps, lagoons, streams, rivers, and zapales (freshwater swamp forest of The Mojana). It is part of the region known as the “Momposina Depression” that is characterized by the flood pulses of the rivers Magdalena, Cauca and San Jorge. In addition, this region is a candidate for World Heritage site (UNESCO 2020) and includes the marsh of Ayapel a Wetland of International Importance (Minambiente 2018). According to the latest municipal measurement to the Multidimensional Poverty Index, this region is also one of the poorest in Colombia (DANE 2015) and its alluvial valleys represent one of the country’s highest levels of anthropic impact according to the first spatiotemporal evaluation of human pressure in Colombia (Correa Ayram et al. 2020).

FIGURE 1. Map showing location of The Mojana and its wetlands prepared by Paola IsaacsCubides. The inset map on the right shows the location of The Mojana in Colombia and the location of Colombia in the Western Hemisphere. The enlarged map on the left shows rivers (Drenaje Doble), marshes (Cienagas - open permanent wetlands), and freshwater swamp forests (Zapales) along with the municipality boundary and the limits of The Mojana. (Source: The Biological Research Institute Alexander von Humboldt, Bogotá.)

1 Instituto de Investigación de Recursos Biológicos Alexander von Humboldt. Corresponding author contact: rayazo@humboldt.org.co. 2 Corporación Paisajes Rurales (Colombia) 3 Universidad de Córdoba (Colombia) 4 Ministry of Environment and Sustainable Development (Colombia) 5 UNDP - United Nations Development Programme (Colombia) Wetland Science & Practice October 2020 323


FIGURE 2. Fisherman in the marsh of Palo (Pescador en Ciénaga de Palo) by Klaudia Cárdenas. This photo received the Best Wetland Photo awarded by Society of Wetland Scientists at the I Congreso Internacional y II Congreso Nacional de Ríos y Humedales. Colombia, 2017.

rescuing the cultural heritage of being a mojanero (an amphibious culture with behaviors, beliefs, and practices related to the management of land and water resources in The Mojana), and monitoring with citizen science (Ayazo Toscano et al. 2020). As a result, between 2016 and 2019, it was possible to implement a socio-ecological rehabilitation through a participatory process that facilitated the recovery of ecological connectivity for 4,822 hectares of this floodplain. To date, these processes of social transformation and ecological rehabilitation are expanding the most successful interventions piloted through the project to an additional eleven municipalities in The Mojana Region (UNDP Colombia 2020). n REFERENCES

Colombia’s Ministry of Environment and Sustainable Development, through a project funded by the Kyoto Protocol’s Adaptation Fund and in partnership with the United Nations Development Programme Colombia, developed a holistic view for climate change and disaster risk management for the region. This included a plan for the rehabilitation of degraded wetland ecosystems and ecosystems-based livelihoods as integral factors of the system necessary for both its operation and the well-being of the locals (Figure 2). In this context, the most feasible strategy to guarantee the success and sustainability of this rehabilitation work was to design a plan based on understanding and recognizing the interdependence of the region’s inhabitants and natural ecosystems (Jaramillo Villa et al. 2018). Through this initiative, the von Humboldt Institute in association with the University of Córdoba and the Rural Landscapes Corporation worked with landowners, communities, local and regional governments to implement and monitor different social-ecological rehabilitation strategies to enhance connectivity of wetland ecosystems. To account for the extremes of droughts and floods in the region, activities focused on the household level by diversifying crops in home gardens, rehabilitating the floodplain landscape,

324 Wetland Science & Practice October 2020

Ayazo Toscano, R., W. Ramírez, and Ú. Jaramillo Villa (Editors.). In Press. Territorios Anfibios en Transición. Rehabilitación Socioecológica de Humedales. Instituto de Investigación de Recursos Biológicos Alexander von Humboldt. Bogotá, D.C., Colombia. Correa Ayram, C.A., A. Etter, J. Díaz-Timoté, S. Rodríguez Buriticá, W. Ramírez, and G. Corzo. 2020. Spatiotemporal evaluation of the human footprint in Colombia: Four decades of anthropic impact in highly biodiverse ecosystems. Ecological Indicators 117: 106630. https://doi. org/10.1016/j.ecolind.2020.106630 DANE. 2015. Pobreza Monetaria y Multidimensional: Principales Resultados 2014, Colombia. Jaramillo Villa, Ú., C. Cárdenas, R. Ayazo Toscano, W. Vargas, N. Gómez, J.C. Linares, M. Carillo, A. Martínez, and W. Ramírez. 2018. Recuperar modos de vida, para rehabilitar ecosistemas: rehabilitación del socio-ecosistema anfibio en la Mojana. In: L.A. Moreno, G.I. Andrade, and M.F. Gómez (Editors). 2019. Biodiversidad 2018. Estado y tendencias de la biodiversidad continental de Colombia. Instituto de Investigación de Recursos Biológicos Alexander von Humboldt. Bogotá, D.C., Colombia. http://reporte.humboldt.org.co/biodiversidad/2018/ cap4/404/#seccion13 Minambiente. 2018. Decreto N° 356. Diario Oficial de la República de Colombia, Bogotá, Colombia, 22 de febrero de 2018 United Nations Development Programme Colombia (UNDP Colombia). 2020. Mojana: Climate and Life. Retrieved from United Nations Development Programme. https://tinyurl.com/y9unbsut UNESCO. 2020. Pre-Hispanic Hydraulic System of the San Jorge River. https://whc.unesco.org/en/tentativelists/5764/


Guardians of Wetlands (Los Guardianes de los Humedales): Young Peruvians Committed to Wetlands Héctor Aponte1

etland conservation is an arduous task; that is why After an arduous and challenging selection process, young people need to get closely involve in its sustain30 guardians were selected for the 2020 GDH program. ability. The Universidad Científica del Sur, one of the Peruvian Between January and March 2020, the guardians received universities that have promoted the study of wetlands in recent training on pre-Inca cultures and coastal wetlands, its funcyears, proposed an alternative that involves young university tioning and characteristics. Also, they had the opportunity students. The Guardianes de los Humedales (Guardians of to do some handicrafts from marshes plants, using tradiWetlands, GDH) program seeks to introduce young people to tional patterns. Likewise, they participated in field workenvironmental research and education related to the protection shops on the identification of native birds and plants as well of coastal wetlands. The study of wetlands is a cross-cutting as they received training to identify the threats and impacts area of ​​interest; it is multi-disciplinary involving lines of to the wetlands visited. research from several university departments such as environOnce trained, the guardians participated in two diffusion activities, situated in the center and south of Lima. mental engineering, agribusiness, tourism, and biology. This program allows us to promote environmental research and These activities were focused on children, where we education from different approaches such as tourism, biodiexplained the importance of wetland conservation through versity management, business studies, evaluation of medicinal different play and learning strategies. Nowadays, the guardplants and land management in Peruvian coastal wetlands. ians are preparing virtual youth awareness activities for The GDH project provides different activities to the the coming months of 2020, as well as meetings with other students. For example, students receive courses and carry young people committed to wetlands (such as the Youth out fieldwork, which allows them to learn about biodiverEngaged in Wetlands initiative). We hoped that as a result sity, culture and research in wetlands. The workshops inof the GDH program, these students would be empowered, clude training in dissemination techniques and awarenessand the network of professionals supporting the protection raising on the protection of coastal wetlands. Therefore, we of wetlands will be strengthened. n can reach a genuinely environmental awareness: FIGURE 1. Guardian of Wetland working with local youth, stimulating an interest in knowledge and motivation. wetlands and their conservation. (Photo by Héctor Aponte.) The GDH program is committed to the training of young people who can play a role as disseminating intermediaries of knowledge, and research and conservation agents of the Peruvian coastal wetlands. It is not possible to achieve environmental awareness without knowledge that sustains the motivation for conservation and that is why we consider the training part as the basis for this project. This program is led by three wetland scientists -Gustavo Lértora, Dámaso W. Ramírez, and Héctor Aponte - who have the support of the Universidad Científica del Sur (Perú) for the execution of the program. It is the second time that this program has been carried out and we are pleased to have the support of the Humedales Costeros (humedalescosteror.org) initiative this time.

W

1 Professional Wetland Scientist. Associate Researcher, Universidad Científica del Sur, Lima-Perú; haponte@cientifica.edu.pe. Wetland Science & Practice October 2020 325


We Are Wetlands (@Somos_humedales) Karina Paz Arteaga1

ince childhood I’ve been concerned about wetlands and their protection. I began my journey when I discovered the Porvenir bay wetland in my hometown (Porvenir Tierra del Fuego in the southernmost region of Chile) where migratory flamingos resided in winte. At the age of 14 I began to study and disseminate scientific information about the Laguna de Los Cisnes wetland located in the same region. By then I knew that I wanted to become a biologist and dedicate my life to wetland conservation. I stopped worrying about wetlands and to take care of them, working on wetland conservation initiatives, highlighting their key role in the conservation of species and their importance to people. Somos_humedales is a personal project reflected through social networks (currently via Facebook and Instagram) designedto provide general knowledge and technical information on wetlands in open access forum (free access) for everyone in a simple language (easy to understand). As such communication and education about wetlands are the main goals of my initiative. The three objectives of this project are: 1) bring the community closer (make aware) to projects that are being developed for the care and protection of wetlands, as well as provide news from reliable sources about conservation, 2) build a connection space (story) of experiences from different territories, with the aim of

S

1 Author contact: karinapaz.medioambiente@gmail.com 326 Wetland Science & Practice October 2020

promoting and facilitating the work of others learning about their experiences and thereby expanding one’s knowledge, and 3) communicate in simple terms technical knowledge about the environment and legal issues so that citizens can decide how to act and what to do to contribute to the care of nature, especially wetlands. For the communication objective, a series of initiatives have been developed that promote the care of wetlands from different organizations, with a citizen vision. The empowerment and identification of the community with wetlands and their biodiversity are fundamental to local wetland protection and make it to highlight projects that have been developed in the region of Los Ríos Chile, especially in cities such as Futrono and Los Lagos with a powerful educational focus. “Somos_humedales” also actively participates in the National Network of Wetlands of Chile. Although this initiative is born from a personal concern and professional interest, it would not be possible without cooperation from people who love wetlands and who have the spirit and conviction support wetland conservation. Collectively, we are part of nature “We are wetlands” (“Somos humedales”). My hope is that, in the near future, “Somos humedales” will become an organization to promote wetland conservation in Chile. n


Urban Wetlands Interactive Platform Camila Teutsch Barros1

any countries in South America and throughout the frame settings. Users can also download the map as a png world still lack the financial and human resources to image, and export data as kml or GeoJSON files for further deliver sound baselines and implement monitoring programs processing through GIS software. to support wetland conservation and management. This exAnother key feature of the platform is the ‘Theme Map’ plains why many wetlands are not even listed within official section. Theme Maps work on the same logic as the Geninformation sources around the world, which makes them eral Map, but also include the option of drawing lines and particularly vulnerable, especially in contexts of rapid urban polygons, and are entirely administered by a particular user growth. This reality is what inspired the development of the or group of users to serve a particular purpose. Urban Wetlands Interactive Platform (Figure 1). The Urban Last but not least, the Platform includes a ‘Campaigns’ Wetlands Interactive Platform is an online participatory map- section, which are based on Theme Maps that are made ping tool that helps gather local information about wetlands public for a certain period of time so as to involve the comand other natural assets world-wide, in order to support munity in participatory processes, citizen science projects water-sensitive urban planning and decision-making. or such. Active Campaigns are showcased in the Platform’s The Platform is not only a cost-effective tool to identify homepage, and Campaign leaders get a Communications Kit which include graphics to help them spread the word wetlands and gather information about them, but also to through social networks. dynamically build on the local knowledge about them and The Platform was developed by Patagua in collaborato share ideas for their protection, restoration and sustaintion with independent researcher Francisco Vasquez. So far, able management. In addition to this, the Platform helps to it is being successfully used in Chile and Colombia. It is connect people and organizations who share an interest on currently available in Spanish only, but there should be an wetlands. It fosters collaboration among them and engages English version coming out in the short term. Learn more local communities in environmental protection. by visiting www.humedalesurbanos.com or contact the The main feature of the Platform is a ‘General Map’ to developers at hola@humedalesurbanos.com. n which registered users can add data points on five different themes (each of which is represented by a color): 1) Initiatives, FIGURE 1. Themes that can be displayed on the interactive platform. 2) Problems, 3) Public Use, 4) Historical Data, and 5) Location. Each point allows users to add a short description as well as a picture. All the information in the General Map is public and free for everyone to explore and comment on. The Platform was developed as an intuitive tool that allows everyone to contribute, no advanced technical skills required. Both street map and satellite view are available as basemaps, and registered users can also save their preferred zoom and

M

1 Executive Director, Patagua; camila@patagua.cl. Wetland Science & Practice October 2020 327


WETLANDS IN THE NEWS

L

isted below are some links to some random news articles that may be of interest. Links from past issues can be accessed on the SWS website: http://sws.org/Sample-Content/wetlands-in-the-news.html. Members are encouraged to send links to articles about wetlands in their local area. Please send the links to WSP Editor at ralphtiner83@gmail.com and reference “Wetlands in the News” in the subject box. Thanks for your cooperation. n Huntington Beach wetlands continue to expand, following decades of degradation https://www.ocregister.com/2020/10/16/huntington-beach-wetlandscontinue-to-expand-following-decades-of-degradation/

Brazil revokes mangrove protections, weakening another ecosystem key to curbing climate change https://www.cnn.com/2020/09/29/americas/brazil-revokes-mangrovesprotection-climate-intl/index.html

The World’s Largest Tropical Wetland Has Become an Inferno https://www.nytimes.com/interactive/2020/10/13/climate/pantanalbrazil-fires.html

Invasive sea lampreys in Great Lakes, and the lake trout they prey on, puzzle scientists https://phys.org/news/2020-09-invasive-sea-lampreys-great-lakes.html

Why Scientists Made Venus Flytraps That Glow https://www.nytimes.com/2020/10/12/science/venus-flytraps-close.html

USDA Awards $5 Million to Support Wetland Mitigation Banking https://www.nrcs.usda.gov/wps/portal/nrcs/detail/national/newsroom/ releases/?cid=NRCSEPRD1660022

To protect nature, bring down the walls of fortress conservation https://www.aljazeera.com/opinions/2020/10/12/to-protect-nature-bringdown-the-walls-of-fortress-conservation/ Florida seeks to take over federal wetland permits https://www.naplesnews.com/story/news/environment/2020/10/12/ florida-seeks-take-over-federal-wetland-permits/3627492001/ Why trout need wetlands https://www.wisconsinwetlands.org/updates/why-trout-need-wetlands/ What’s Green, Soggy and Fights Climate Change? https://www.nytimes.com/2020/10/09/climate/peat-climate-change.html Droughts are threatening global wetlands https://phys.org/news/2020-10-droughts-threatening-global-wetlands.html How to reverse global wildlife declines by 2050 https://theconversation.com/how-to-reverse-global-wildlife-declinesby-2050-146041 Elusive eastern black rail threatened by rising sea levels https://apnews.com/article/habitat-destruction-wildlife-climate-changerising-sea-levels-climate-5a8ea861445582c2625d93ba82069c70 Mercury on the Rise https://wrri.ncsu.edu/blog/2019/11/mercury-on-the-rise/ 40 Percent of the Amazon Is on the Brink of Becoming Savanna https://earther.gizmodo.com/40-percent-of-the-amazon-is-on-the-brinkof-transitioni-1845274803 USDA Seeks New Partnerships to Safeguard, Restore Wetland Ecosystems https://www.nrcs.usda.gov/wps/portal/nrcs/detail/national/newsroom/ releases/?cid=NRCSEPRD1664617 How to Revolutionize Biodiversity Conservation in the U.S. https://www.scientificamerican.com/article/how-to-revolutionize-biodiversity-conservation-in-the-u-s/ MR-GO closure improving environment, but more wetlands, swamp restoration needed https://www.nola.com/news/environment/article_06d67378-0423-11ebbcbf-5f2aad29d796.html 40 Percent of World's Plants at Risk of Extinction https://www.ecowatch.com/plants-biodiversity-extinction-2647868027.html New Study Shows a Vicious Circle of Climate Change Building on Thickening Layers of Warm Ocean Water https://insideclimatenews.org/news/28092020/ocean-stratificationclimate-change

Blue Origin to fill wetlands for rocket test site https://www.floridatoday.com/story/news/local/environment/2020/09/29/ blue-origin-fill-wetlands-rocket-test-site/3559508001/ Why Nova Scotia wants to poison a lake to kill off invasive species https://www.cbc.ca/news/canada/nova-scotia/nova-scotia-lake-poisoninvasive-species-1.5739446 A Watershed Study for Wetland Restoration https://www.newswise.com/articles/a-watershed-study-for-wetlandrestoration Study finds spreading ghost forests on NC coast may contribute to climate change https://phys.org/news/2020-09-ghost-forests-nc-coast-contribute.html The World's Marvellously Freaky Carnivorous Plants Are in More Trouble Than We Knew https://www.sciencealert.com/the-world-s-marvellously-freaky-carnivorous-plants-are-in-dire-need-of-our-help Two new species of wetland plant discovered from Western Ghats https://indianexpress.com/article/india/two-new-species-of-wetlandplant-discovered-from-western-ghats-6619464/ Wetland Silviculture & Water Tables https://www.srs.fs.usda.gov/compass/2020/01/30/wetland-silviculturewater-tables/ Florida Gulf Coast University unveils new ultrasonic technology to fight algae blooms https://www.fox4now.com/news/protecting-paradise/florida-gulf-coastuniversity-unveils-new-ultrasonic-technology-to-fight-algae-blooms Extraordinary Fires Char the Pantanal, a Vast Floodplain in South America https://scitechdaily.com/extraordinary-fires-char-the-pantanal-a-vastfloodplain-in-south-america/ Brazilian wetlands fires started by humans and worsened by drought https://www.theguardian.com/world/2020/sep/18/brazilian-wetlandsfires-started-by-humans-and-worsened-by-drought Army Corps launches Everglades, Biscayne Bay restoration plan https://www.miamiherald.com/news/local/environment/article245828505.html New wetland recharge park opens in Ocala https://www.wcjb.com/2020/09/17/new-wetland-recharge-park-opensin-ocala/

The world’s largest wetland is on fire: how can we save the Pantanal? https://blog.ecosia.org/what-is-the-pantanal-and-why-is-it-burning/

‘The warning lights are flashing.’ Report finds nations failing to protect biodiversity https://www.sciencemag.org/news/2020/09/warning-lights-are-flashingreport-finds-nations-failing-protect-biodiversity

'The Blob': Low-oxygen water killing lobsters, fish in Cape Cod Bay https://www.capecodtimes.com/news/20200928/the-blob-low-oxygenwater-killing-lobsters-fish-in-cape-cod-bay

Handing federal wetlands permitting to FL DEP is an idea that’s all wet https://www.floridaphoenix.com/2020/09/17/handing-federal-wetlandspermitting-to-fl-dep-is-an-idea-thats-all-wet/

328 Wetland Science & Practice October 2020


Fires in Pantanal, world's largest tropical wetlands, 'triple' in 2020 https://www.bbc.com/news/world-latin-america-53500288 Portugal's airport plans threaten wetlands https://science.sciencemag.org/content/369/6510/1440.1 Australian stinging trees release spider-like venom that can hurt for weeks https://www.cnn.com/2020/09/17/asia/australia-venomous-stingingtrees-scn-scli-intl/index.html Soggy coastal soils? Here’s why ecologists love them https://www.sciencenewsforstudents.org/article/wetlands-coastal-soilsecology-flooding-storm-surge-climate-change Long-term change in habitat and vegetation in an ungrazed, estuarine salt marsh: Man-made foreland compared to young marsh development https://www.sciencedirect.com/science/article/abs/pii/ S0272771419302537 Conservationists Split Over Poseidon Desal Project’s Potential to Help Bolsa Chica Wetlands Voice of OC https://voiceofoc.org/2020/09/conservationists-split-over-poseidondesal-projects-potential-to-help-bolsa-chica-wetlands/ Birds are mysteriously dying in New Mexico in 'frightening' numbers https://www.lcsun-news.com/story/news/2020/09/12/mass-deaths-migratory-birds-new-mexico-environment/5780282002/ Battle on to save Brazil's tropical wetlands from flames https://apnews.com/35368749c3619220d79a2ecfa9ab1e3b Trump Administration Announces more than $130 Million in PublicPrivate Funding for Wetland Conservation Projects https://www.mychesco.com/a/news/national/trump-administration-announces-more-than-130-million-in-public-private-funding-for-wetlandconservation-projects/ Buzzy’s Ranch wetlands restoration seeks grant to buy water rights https://sierranevadaally.org/2020/09/09/buzzys-ranch-wetlands-restoration-looks-for-grant-to-buy-water-rights/ World's wildlife populations in devastating decline warns WWF report https://www.cnn.com/2020/09/09/world/wwf-report-species-declineclimate-scn-intl-scli/index.html More than half of government environmental scientists say their work has been suppressed https://www.abc.net.au/news/science/2020-09-09/environment-scientistscensored-suppressed-data/12643824 Wildlife in 'catastrophic decline' due to human destruction, scientists warn https://www.bbc.com/news/science-environment-54091048 Bending the Curve of Biodiversity Loss: Ambitious Conservation and Restoration Efforts Required https://scitechdaily.com/bending-the-curve-of-biodiversity-loss-ambitious-conservation-and-restoration-efforts-required/ As the Arctic thaws, Indigenous Alaskans demand a voice in climate change research https://www.sciencemag.org/news/2020/09/arctic-thaws-indigenousalaskans-demand-voice-climate-change-research How artificial salt marshes can help in the fight against rising seas https://www.theguardian.com/environment/2020/sep/09/how-artificialsalt-marshes-can-help-in-the-fight-against-rising-seas-aoe With Baylands under flood threat, Palo Alto explores projects to address sea level rise https://www.paloaltoonline.com/news/2020/09/07/with-baylands-underflood-threat-palo-alto-explores-projects-to-address-sea-level-rise Why plants in wetlands are highly productive https://phys.org/news/2020-09-wetlands-highly-productive.html Wetland area does not get Greene County stamp of approval, land trust makes plans for another proposal https://www.daytondailynews.com/news/wetland-area-does-not-getgreene-county-stamp-of-approval-land-trust-makes-plans-for-anotherproposal/HV5ON7B34ZCOBN5GVCXRTCPUMI/

Pitkin County launches project to restore ancient wetland at North Star Preserve near Aspen https://www.9news.com/article/life/style/colorado-guide/pitkincounty-project-restore-ancient-wetland-north-star-preserveaspen/73-78622b09-815a-48ff-938f-999d67771d77 New floating wetlands in Winton Woods pond hold aquatic plants https://www.cincinnati.com/story/news/2020/09/05/new-floating-wetlands-winton-woods-pond-hold-aquatic-plants/5730368002/ Invasive green crab species spotted in Padilla Bay, Anacortes https://mynorthwest.com/2139258/invasive-green-crabs-padilla-bay/ Invasive aquatic plant found in 4 Michigan inland lakes https://www.wnem.com/news/invasive-aquatic-plant-found-in-4-michigan-inland-lakes/article_9aa160ae-ef97-11ea-978a-3381cd594084.html Brazil Pantanal Scorched by Fires https://www.nytimes.com/2020/09/04/world/americas/brazil-wetlandsfires-pantanal.html Court Overturns Administration Efforts to Weaken the Migratory Bird Treaty Act https://abcbirds.org/article/court-upholds-mbta-2020/ Quakertown swamp gets Wetland of Distinction title | Southeastern Pennsylvania https://www.wfmz.com/news/area/southeastern-pa/quakertown-swampgets-wetland-of-distinction-title/article_f6c60f74-e7ea-11ea-838d4707c0188c10.html NASA Research Reveals the True Causes of Sea Level Rise Since 1900 https://scitechdaily.com/nasa-research-reveals-the-true-causes-of-sealevel-rise-since-1900/ Human-Nature Connectivity: Wetlands Within Sustainable Futures https://link.springer.com/chapter/10.1007/978-3-030-41306-4_6 Hamoun wetlands restoration project started https://www.tehrantimes.com/news/451584/Hamoun-wetlands-restoration-project-started Inverleith Park Pond - Water Gems https://www.watergems.co.uk/portfolio_page/inverleith-park-pond/ Massachusetts grants $798,500 for river, wetland restoration projects https://www.masslive.com/news/2020/08/massachusetts-grants798500-in-river-wetland-restoration-projects.html BCCC & Sound Rivers to start construction on new wetlands https://www.witn.com/2020/08/22/bccc-sound-rivers-to-start-construction-on-new-wetlands/ Potsdam professor turns to beetles in fight against invasive purple loosetrife https://www.newyorkupstate.com/adirondacks/2020/08/potsdam-professor-turns-to-beetles-in-fight-against-invasive-purple-loosetrife.html VEWH - Improving plants in Lower Barwon wetlands https://www.vewh.vic.gov.au/news-and-publications/stories/improvingplants-in-lower-barwon-wetlands Brazil fires threaten world's largest wetland https://www.reuters.com/article/us-brazil-environment-fires-idUSKCN25E2FC Watch hummingbirds ‘dance’ through waterfalls https://www.sciencemag.org/news/2020/08/watch-hummingbirds-dancethrough-waterfalls The sea otter rescue plan that worked too well https://www.bbc.com/future/article/20200818-the-canadian-sea-otterrescue-plan-that-worked-too-well Carnivorous Plants of Delaware's Wetlands: sundew, bladderworts & more https://wmap.blogs.delaware.gov/2020/03/05/wetland-carnivorousplants-nothing-to-be-afraid-of/ Under the radar rollback of stream and wetland protections https://thehill.com/opinion/energy-environment/512499-under-the-radarrollback-of-stream-and-wetland-protections

Wetland Science & Practice October 2020 329


Greenland's ice sheet has melted to a point of no return, study finds https://www.cnn.com/2020/08/14/weather/greenland-ice-sheet/index.html Concrete to be removed to create wetlands in East Bay https://www.eastbaytimes.com/old-aircraft-taxiway-in-alameda-to-beconverted-to-wetlands-park 'It overtakes everything:' Invasive aquatic plant threatens to overrun Potomac watershed https://wjla.com/news/local/invasive-plant-species-potomac-watershed Scientists decry federal rule that removes protection from 'unconnected' streams and wetlands https://phys.org/news/2020-08-scientists-decry-federal-unconnectedstreams.html High-latitude Climate Change - Alaska Nature and Science (U.S. National Park Service) https://www.nps.gov/subjects/aknatureandscience/hi-latclimatechange.htm Dead sturgeon found on Lake Michigan beaches https://www.michiganradio.org/post/dead-sturgeon-found-lake-michigan-beaches We mapped the world's frozen peatlands – what we found was very worrying https://theconversation.com/we-mapped-the-worlds-frozen-peatlandswhat-we-found-was-very-worrying-144235 Half Of Wisconsin’s Wetlands To Lose Federal Protection, Nature Conservancy Estimates https://www.wxpr.org/post/half-wisconsin-s-wetlands-lose-federal-protection-nature-conservancy-estimates#stream/0 A century ago, it was a Florida landscape shrub. Now, it's a pest plant that keeps spreading. https://phys.org/news/2020-08-century-florida-landscape-shrub-pest.html

Fridays On The Farm: Restoring Wetlands and Creating Habitat https://www.farmers.gov/connect/blog/conservation/fridays-farm-restoring-wetlands-and-creating-habitat Climate change: UK peat emissions could cancel forest benefits https://www.bbc.com/news/science-environment-53684047 Wind farms built on carbon-rich peat bogs lose their ability to fight climate change https://theconversation.com/wind-farms-built-on-carbon-rich-peat-bogslose-their-ability-to-fight-climate-change-143551 Articles on international activities in wetland and nature conservation https://news.mongabay.com/author/genevieve/ New wetlands open to the public https://www.dailystandard.com/archive/2020-08-03/stories/41000/newwetlands-open-to-the-public Peat takes millennia to generate, and bogs store 10 times more carbon than forests — using it in gardening is madness https://www.countrylife.co.uk/nature/peat-takes-millennia-to-generateand-bogs-store-10-times-more-carbon-than-forests-using-it-in-gardening-is-madness-214851 Ancient bones in disturbed peat bogs are rotting away, alarming archaeologists https://www.sciencemag.org/news/2020/07/ancient-bones-disturbedpeat-bogs-are-rotting-away-alarming-archaeologists Smaller habitats worse than expected for biodiversity https://phys.org/news/2020-07-smaller-habitats-worse-biodiversity.html Kabartal Wetland https://definearth.com/2020/04/05/kabartal-wetland/

UNEP supports project to restore peatlands in Indonesia https://www.unenvironment.org/news-and-stories/story/unep-supportsproject-restore-peatlands-indonesia

'A win for the Everglades': 5,000 pythons removed in state-sponsored capture program https://www.msn.com/en-us/news/us/a-win-for-the-everglades5-000-pythons-removed-in-state-sponsored-capture-program/arBB17lBQM?li=BBnbcA1

Wetland revival https://www.iowafarmbureau.com/Article/Wetland-revival

Wetland Wildlife of Arkansas http://www.naturalheritage.com/blog/wetland-wildlife

Market St. development on wetland overcomes opposition, heads to Wilmington council https://portcitydaily.com/local-news/2020/08/10/market-st-developmenton-wetland-overcomes-opposition-heads-to-wilmington-council/

Ireland's Peatlands - The National Trust for Ireland https://www.antaisce.org/issues/irelands-peatlands

Citrus Flavoring Is Weaponized Against Insect-Borne Diseases https://www.nytimes.com/2020/08/10/health/tick-mosquito-repellantnootkatone.html Abandoned Fiberglass Boats are Polluting the Marine Environment https://www.maritime-executive.com/editorials/abandoned-fiberglassboats-are-polluting-the-marine-environment Toxic Chemicals From Fossil Fuels Are Poisoning East Coast Dolphins and Whales, Study Finds https://www.ecowatch.com/toxic-chemicals-dolphins-whalesus-2646906589.html Plants cropping up in lost Michigan lakes where dams failed https://news.yahoo.com/plants-cropping-lost-michiganlakes-130604416.html

330 Wetland Science & Practice October 2020

New wetland installed in Putnam County https://www.limaohio.com/news/419468/new-wetland-installed-inputnam-county Two new bird species discovered nesting in restored Arkansas wetlands https://www.kark.com/news/local-news/two-new-bird-species-discovered-nesting-in-restored-arkansas-wetlands/ Plan to dissolve Peatland Restoration Agency raised concerns https://www.thejakartapost.com/news/2020/07/16/plan-to-dissolve-peatland-restoration-agency-raised-concerns.html Global methane emissions soar to record high https://phys.org/news/2020-07-global-methane-emissions-soar-high.html


WETLAND BOOKSHELF

oug Wilcox, former editor of Wetlands, has edited and produced a book that provides a review of the history of wetland science from the experiences of today’s wetland scientists. The book entitled “History of Wetland Science: A Perspective from Wetland Leaders” was published in July 2020 and is available from Amazon (https://www. amazon.com/History-Wetland-Science-Perspectives-Leaders/dp/B08DC6GXDM). In the Preface, Doug reveals that his inspiration for the book came from the fact that many of his old wetland friends were retiring, retired and a few had passed and he wanted to record some of their experiences. He felt it would be a meaningful exercise to have wetland scientists who have made significant contributions to wetland science, management, or conservation provide short autobiographies that would offer presentday and future students of wetland science some perspective on how early wetland scientists became involved in wetlands and the challenges they faced in their careers. Students as well as today’s wetland scientists should find the chapters interesting, inspiring, and sometimes humorous as contributors offer a personal glimpse of their life with wetlands.

D

BOOKS

• History of Wetland Science: A Perspective from Wetland Leaders https://www.amazon.com/History-Wetland-Science-Perspectives-Leaders/dp/B08DC6GXDM • An Introduction to the Aquatic Insects of North America (5th Edition) https://he.kendallhunt.com/product/introduction-aquatic-insects-north-america • Wading Right In: Discovering the Nature of Wetlands https://press.uchicago.edu/ucp/books/book/chicago/W/ bo28183520.html • Sedges of Maine https://umaine.edu/umpress/books-in-print/ • Sedges and Rushes of Minnesota https://www.upress.umn. edu/book-division/books/sedges-and-rushes-of-minnesota • Wetland & Stream Rapid Assessments: Development, Validation, and Application https://www.elsevier.com/ books/wetland-and-stream-rapid-assessments/dorney/978-0-12-805091-0 • Eager: The Surprising Secret Life of Beavers and Why They Matter https://www.chelseagreen.com/product/eager/ • Wetland Indicators – A Guide to Wetland Formation, Identification, Delineation, Classification, and Mapping https://www.crcpress.com/Wetland-Indicators-A-Guide-toWetland-Identification-Delineation-Classification/Tiner/p/ book/9781439853696

Although the list of participating scientists is huge (71 contributors) and contains the names of many familiar wetland scientists, it is not exhaustive as others were unable to contribute for various reasons. Nonetheless, the book offers a unique frame of reference for presenting the history of wetland science by having today’s wetland scientists provide short stories of their introduction to wetlands and a reflection on their careers. For the latest news on wetlands and related topics, readers are referred to the Association of State Wetland Managers website: https://www.aswm.org/. Their “Wetland News Digest” section include links to newspaper articles that should be of interest: https://www.aswm.org/news/ wetland-breaking-news. Please help us add new books to this listing. If your agency, organization, or institution has a website where wetland information can be accessed, please send the information to the Editor of Wetland Science & Practice at ralphtiner83@gmail. com. Your cooperation is appreciated. n

• Wetland Soils: Genesis, Hydrology, Landscapes, and Classification https://www.crcpress.com/Wetland-Soils-Genesis-Hydrology-Landscapes-and-Classification/VepraskasRichardson-Vepraskas-Craft/9781566704847 • Creating and Restoring Wetlands: From Theory to Practice http://store.elsevier.com/Creating-and-Restoring-Wetlands/ Christopher-Craft/isbn-9780124072329/ • Salt Marsh Secrets. Who uncovered them and how? http://trnerr.org/SaltMarshSecrets/ • Remote Sensing of Wetlands: Applications and Advances. https://www.crcpress.com/product/isbn/9781482237351 • Wetlands (5th Edition). http://www.wiley.com/WileyCDA/ WileyTitle/productCd-1118676823.html • Black Swan Lake – Life of a Wetland http://press.uchicago. edu/ucp/books/book/distributed/B/bo15564698.html • Coastal Wetlands of the World: Geology, Ecology, Distribution and Applications http://www.cambridge.org/ us/academic/subjects/earth-and-environmental-science/ environmental-science/coastal-wetlands-world-geologyecology-distribution-and-applications • Florida’s Wetlands https://www.amazon.com/FloridasWetlands-Natural-Ecosystems-Species/dp/1561646873/ ref=sr_1_4?ie=UTF8&qid=1518650552&sr=84&keywords=wetland+books Wetland Science & Practice October 2020 331


• Mid-Atlantic Freshwater Wetlands: Science, Management, Policy, and Practice http://www.springer.com/environment/ aquatic+sciences/book/978-1-4614-5595-0 • The Atchafalaya River Basin: History and Ecology of an American Wetland http://www.tamupress.com/product/ Atchafalaya-River-Basin,7733.aspx • Tidal Wetlands Primer: An Introduction to their Ecology, Natural History, Status and Conservation https://www. umass.edu/umpress/title/tidal-wetlands-primer • Wetland Landscape Characterization: Practical Tools, Methods, and Approaches for Landscape Ecology http:// www.crcpress.com/product/isbn/9781466503762 • Wetland Techniques (3 volumes) http://www.springer.com/ life+sciences/ecology/book/978-94-007-6859-8 • Wildflowers and Other Plants of Iowa Wetlands https://www.uipress.uiowa.edu/books/2015-spring/wildflowers-and-other-plants-iowa-wetlands.htm • Wetland Restoration: A Handbook for New Zealand Freshwater Systems https://www.landcareresearch.co.nz/ publications/books/wetlands-handbook • Wetland Ecosystems https://www.wiley.com/en-us/ Wetland+Ecosystems-p-9780470286302 • Constructed Wetlands and Sustainable Development https://www.routledge.com/Constructed-Wetlands-and-Sustainable-Development/Austin-Yu/p/ book/9781138908994

332 Wetland Science & Practice October 2020

ONLINE SOURCES OF WETLAND INFORMATION The following is a listing of some government agencies and environmental organizations that provide online information on wetlands and where their publications on wetlands may be accessed. • U.S. Army Corps of Engineers, Waterways Experiment Station, Environmental Laboratory https://www.erdc. usace.army.mil/Locations/EL.aspx • U.S. Army Corps of Engineers, National Wetland Plants Database http://wetland-plants.usace.army.mil/nwpl_static/v34/home/home.html • U.S. Environmental Protection Agency https://www.epa. gov/wetlands • U.S. Fish and Wildlife Service, National Wetlands Inventory https://fws.gov/wetlands/ • U.S. Geological Survey, Wetland and Aquatic Research Center https://www.usgs.gov/centers/wetland-and-aquaticresearch-center-warc • U.S. Geological Survey, Northern Prairie Wildlife Research Center https://www.usgs.gov/centers/npwrc • U.S. Geological Survey, Patuxent Wildlife Research Center https://www.usgs.gov/centers/pwrc • National Oceanic and Atmospheric Administration, Office of Coastal Management https://coast.noaa.gov/ • U.S.D.A. Natural Resources Conservation Service, Hydric Soils https://www.nrcs.usda.gov/wps/portal/nrcs/main/ soils/use/hydric/ • Association of State Wetland Managers https://www. aswm.org/


WETLANDS JOURNAL

What’s New in the SWS Journal - Wetlands? The following articles appear in Volume 40, Issue 4 of Wetlands, Journal of the Society of Wetland Scientists. • The Twin Limit Marsh Model: A Non-equilibrium Approach to Predicting Marsh Vegetation on Shorelines and in Floodplains • Denitrification Capacity of Hill Country Wet and Dry Area Soils as Influenced by Dissolved Organic Carbon Concentration and Chemistry • Dissolved Organic Matter Export from Surface Sediments of a New England Salt Marsh • Nutrient Inputs and Hydrology Interact with Plant Functional Type in Affecting Plant Production and Nutrient Contents in a Wet Grassland • Patterns of Denitrification and Methanogenesis Rates from Vernal Pools in a Temperate Forest Driven by Seasonal, Microbial Functional Gene Abundances, and Soil Chemistry • The Difference in Light use Efficiency between an Abandoned Peatland Pasture and an Adjacent Boreal Bog in Western Newfoundland, Canada • A Suitable Method for Assessing Invasibility of Habitats in the Ramsar Sites - an Example of the Southern Part of the Pannonian Plain • Multiple Stressors Influence Salt Marsh Recovery after a Spring Fire at Mugu Lagoon, CA • Impact of Barrier Breaching on Wetland Ecosystems under the Influence of Storm Surge, Sea-Level Rise and Freshwater Discharge • Feedbacks of Alpine Wetlands on the Tibetan Plateau to the Atmosphere • Coastal Marsh Bird Habitat Selection and Responses to Hurricane Sandy • Do Breeding Bird Communities or Conservation Value Differ Among Forested Wetland Types or Ecoregions in Nova Scotia? • The Missing Metric: An Evaluation of Fungal Importance in Wetland Assessments • Foundation Species Loss Affects Leaf Breakdown and Aquatic Invertebrate Resource Use in Black Ash Wetlands • Habitat Suitability Modelling of Benthic Macroinvertebrate Community in Wetlands of Lake Tana Watershed, Northwest Ethiopia • Impact of Peatland Restoration on Soil Microbial Activity and Nematode Communities • Plant Community Establishment in a Coastal Marsh Restored Using Sediment Additions • A Framework for Considering Climate Change Impacts in Project Selection for Deepwater Horizon Restoration Efforts • In-Situ CO2 Partitioning Measurements in a Phragmites australis Wetland: Understanding Carbon Loss through Ecosystem Respiration • Correction to: Morphology of Drained Upland Depressions on the Des Moines Lobe of Iowa • Correction to: “Relationships Between Salinity and ShortTerm Soil Carbon Accumulation Rates from Marsh Types Across a Landscape in the Mississippi River Delta"

Wetland Science & Practice October 2020 333


WSP SUBMISSION GUIDELINES

About Wetland Science & Practice (WSP) etland Science and Practice (WSP) is the SWS quarterly publication aimed at providing information on select SWS activities (technical committee summaries, chapter workshop overview/ abstracts, and SWS-funded student activities), brief summary articles on ongoing or recently completed wetland research, restoration, or management projects or on the general ecology and natural history of wetlands, and highlights of current events. WSP also includes sections listing new publications and research at various institutions, and links to major wetland research facilities, federal agencies, wetland restoration/monitoring sites and wetland mapping sites. The publication also serves as an outlet for commentaries, perspectives and opinions on important developments in wetland science, theory, management and policy. Both invited and unsolicited manuscripts are reviewed by the WSP editor for suitability for publication. Student papers are welcomed. Please see publication guidelines at the end of this issue. Electronic access to Wetland Science and Practice is included in your SWS membership. All issues published, except the the current issue, are available via the internet to the general public. At the San Juan meeting, the SWS Board of Directors voted to approve release of past issues of WSP when a new issue is available to SWS members only. This means that a WSP issue will be available to the public four months after it has been read by SWS members (e.g., the June 2017 issue will be an open access issue in September 2017). Such availability will hopefully stimulate more interest in contributing to the journal. And, we are excited about this opportunity to promote the good work done by our members.

W

HOW YOU CAN HELP If you read something you like in WSP, or that you think someone else would find interesting, be sure to share. Share links to your Facebook, Twitter, Instagram and LinkedIn accounts. Make sure that all your SWS colleagues are checking out our recent issues, and help spread the word about SWS to non-members! Questions? Contact editor Ralph Tiner, PWS Emeritus (ralphtiner83@gmail.com). n 334 Wetland Science & Practice October 2020

WSP Manuscript – General Guidelines LENGTH: Approximately 5,000 words; can be longer if necessary. STYLE: See existing articles from 2014 to more recent years available online at: http://www.sws.org/Publications/wsp-contents.html TEXT: Word document, 12 font, Times New Roman, single-spaced; keep tables and figures separate, although captions can be included in text. For reference citations in text use this format: (Smith 2016; Jones and Whithead 2014; Peterson et al. 2010). FIGURES: Please include color images and photos of subject wetland(s) as WSP is a full-color e-publication. Image size should be less than 1MB – 500KB may work best for this e-publication. REFERENCE CITATION EXAMPLES: • Claus, S., S. Imgraben, K. Brennan, A. Carthey, B. Daly, R. Blakey, E. Turak, and N. Saintilan. 2011. Assessing the extent and condition of wetlands in NSW: Supporting report A – Conceptual framework, Monitoring, evaluation and reporting program, Technical report series, Office of Environment and Heritage, Sydney, Australia. OEH 2011/0727. • Clements, F.E. 1916. Plant Succession: An Analysis of the Development of Vegetation. Carnegie Institution of Washington. Washington D.C. Publication 242. • Clewell, A.F., C. Raymond, C.L. Coultas, W.M. Dennis, and J.P. Kelly. 2009. Spatially narrow wet prairies. Castanea 74: 146-159. • Colburn, E.A. 2004. Vernal Pools: Natural History and Conservation. McDonald & Woodward Publishing Company, Blacksburg, VA. • Cole, C.A. and R.P. Brooks. 2000. Patterns of wetland hydrology in the Ridge and Valley Province, Pennsylvania, USA. Wetlands 20: 438-447. • Cook, E.R., R. Seager, M.A. Cane, and D.W. Stahle. 2007. North American drought: reconstructions, causes, and consequences. Earth-Science Reviews 81: 93-134. • Cooper, D.J. and D.M. Merritt. 2012. Assessing the water needs of riparian and wetland vegetation in the western United States. U.S.D.A., Forest Service, Rocky Mountain Research Station, Ft. Collins, CO. Gen. Tech. Rep. RMRS-GTR-282.


WEB TIP

Resources at your fingertips! For your convenience, SWS has compiled a hefty list of wetland science websites, books, newsletters, government agencies, research centers and more, and saved them to sws.org. Find them on the Related Links page sws.org.

Coming in 2021! ADVERTISING OPPORTUNITIES IN WSP Stay tuned for more information to be released soon!

Wetland Science Practice

&

WSP is the formal voice of the Society of Wetland Scientists. It is a quarterly publication focusing on the news of the SWS and providing important announcements for members and opportunities for wetland scientists, managers, and graduate students to publish brief summaries of their works and conservation initiatives. Topics for articles may include descriptions of threatened wetlands around the globe or the establishment of wetland conservation areas, and summary findings from research or restoration projects. All manuscripts should follow guidelines for authors listed above. All papers published in WSP will be reviewed by the editor for suitability and may be subject to peer review as necessary. Most articles will be published within 3 months of receipt. Letters to the editor are also encouraged, but must be relevant to broad wetland-related topics. All material should be sent electronically to the current editor of WSP. Complaints about SWS policy or personnel should be sent directly to the elected officers of SWS and will not be considered for publication in WSP. n

SOCIETY OF WETLAND SCIENTISTS 1818 Parmenter St., Ste 300, Middleton, WI 53562 (608) 310-7855 www.sws.org

Wetland Science & Practice October 2020 335


Issuu converts static files into: digital portfolios, online yearbooks, online catalogs, digital photo albums and more. Sign up and create your flipbook.